<<

1 Orbital Evidence for Clay and Acidic Sulfate Assemblages on Based on 2 Mineralogical Analogs from Rio Tinto, Spain 3 4 Hannah H. Kaplan1*, Ralph E. Milliken1, David Fernández-Remolar2, Ricardo Amils3, 5 Kevin Robertson1, and Andrew H. Knoll4 6 7 Authors: 8 9 1 Department of Earth, Environmental and Planetary Sciences, Brown University, 10 Providence, RI, 02912 USA. 11 12 2British Geological Survey, Nicker Hill, Keyworth, NG12 5GG, UK. 13 14 3 Centro de Astrobiologia (INTA-CSIC), Ctra Ajalvir km 4, Torrejon de Ardoz, 28850, 15 Spain. 16 17 4 Department of Organismic and Evolutionary Biology, Harvard University, 18 Çambridge, MA, USA. 19 20 *Corresponding author: [email protected]

1 21 Abstract 22 23 Outcrops of hydrated minerals are widespread across the surface of Mars,

24 with clay minerals and sulfates being commonly identified phases. Orbitally-based

25 reflectance spectra are often used to classify these hydrated components in terms of

26 a single mineralogy, although most surfaces likely contain multiple minerals that

27 have the potential to record local geochemical conditions and processes.

28 Reflectance spectra for previously identified deposits in Ius and Melas

29 within the , Mars, exhibit an enigmatic feature with two distinct

30 absorptions between 2.2 – 2.3 µm. This spectral ‘doublet’ feature is proposed to

31 result from a mixture of hydrated minerals, although the identity of the minerals has

32 remained ambiguous. Here we demonstrate that similar spectral doublet features

33 are observed in airborne, field, and laboratory reflectance spectra of rock and

34 sediment samples from Rio Tinto, Spain. Combined visible-near infrared reflectance

35 spectra and X-ray diffraction measurements of these samples reveals that the

36 doublet feature arises from a mixture of Al-phyllosilicate (illite or muscovite) and

37 jarosite. Analyses of orbital data from the Compact Reconnaissance Imaging

38 Spectrometer for Mars (CRISM) shows that the martian spectral equivalents are also

39 consistent with mixtures of Al-phyllosilicates and jarosite, where the Al-

40 phyllosilicate may also include kaolinite and/or halloysite. A case study for a region

41 within demonstrates that the relative proportions of the Al-

42 phyllosilicate(s) and jarosite vary within one stratigraphic unit as well as between

43 stratigraphic units. The former observation suggests that the jarosite may be a

44 diagenetic (authigenic) product and thus indicative of local pH and redox conditions,

2 45 whereas the latter observation may be consistent with variations in sediment flux

46 and/or fluid chemistry during sediment deposition.

47

48 Key points: 49 • A doublet absorption is observed in reflectance spectra of Mars between 2.2 - 50 2.3 µm. 51 • Reflectance spectra at Rio Tinto, Spain also show a doublet due to Al- 52 phyllosilicate and jarosite. 53 • Spectral variation in martian deposits suggests local variation in pH, redox, 54 and/or sediment flux. 55 56 57 Keywords: Mars, surface; Spectroscopy; Infrared observations 58

3 59 1. Introduction

60 Outcrops of sulfate and clay mineral-bearing rocks are widespread across

61 the ancient surface of Mars [Bibring et al., 2006; Murchie et al., 2009; Ehlmann et al.,

62 2011; Carter et al., 2013], indicating that a complex history of water-rock interaction

63 is preserved in the martian rock record. Visible-near infrared (VIS-NIR) reflectance

64 spectra acquired by the CRISM (Compact Reconnaissance Imaging Spectrometer for

65 Mars) and OMEGA (Observatoire pour la Minéralogie, l’Eau, les Glaces et l’Activité)

66 imaging spectrometers have been used to identify clay minerals in ancient and

67 heavily cratered terrains [Poulet et al., 2005; Ehlmann et al., 2011; Carter

68 et al., 2013] and thick deposits of sulfates in younger terrains and basins

69 [Gendrin et al., 2005; Bibring et al., 2006]. In the cases of Terra Meridiani,

70 Crater, and Crater, these minerals have been examined in situ by rover-based

71 payloads, providing detailed local geologic context for some of these orbital

72 detections [e.g. Klingelhofer et al., 2004; Squyres et al., 2004; Arvidson et al., 2008;

73 Wang et al., 2006; Vaniman et al., 2014].

74

75 Many of these orbiter-based detections of hydrated minerals on Mars are

76 discussed in terms of single minerals. That is, the reflectance spectra of these

77 deposits are commonly discussed or interpreted as specific clay, sulfate, or other

78 hydrated minerals (see Ehlmann et al., 2009 for an exception with regards to

79 mineral assemblages at ). Though these deposits are undoubtedly

80 polymineralic, it appears that in many cases a single mineral can be spectrally

81 dominant even at the spatial resolution of instruments such as CRISM (e.g., ~18

4 82 m/pixel). Therefore, those regions on Mars whose reflectance spectra cannot clearly

83 be attributed to a single dominant hydrated mineral are of particular interest as

84 they may record different and more diverse geological processes than are typically

85 recognized from CRISM data.

86

87 Reflectance spectra of deposits and strata within the Valles Marineris reveal

88 a number of locations that fall within this category, and these are the focus of this

89 study. Specifically, CRISM spectra of various deposits within Ius, Melas, and other

90 chasmata exhibit an absorption feature with local minima centered at ~2.21 and

91 ~2.27 μm [Metz et al., 2009; Roach et al., 2010; Weitz et al., 2011] (Figure 1). The

92 spectral shape and properties of this absorption feature, which we will refer to as a

93 spectral ‘doublet’, does not match the spectrum of any single common mineral in

94 existing spectral libraries, suggesting it likely cannot be attributed to a single type of

95 clay, sulfate, or other common hydrated mineral.

96

97 Previous studies have hypothesized that this material represents a mixture

98 of clays, sulfates (possibly including jarosite), and/or poorly crystalline phases

99 [Roach et al., 2010; Tosca et al., 2008b; Noe Dobrea et al., 2011; Weitz et al., 2014],

100 but the exact assemblage remains unknown. We have observed that similar spectral

101 features occur in airborne, field, and lab spectra of rocks and sediments at Rio Tinto,

102 Spain, which has been previously studied as a Mars analog due to the river’s acidity,

103 iron hydrochemistry, and extensive iron oxide deposits [Fernandez-Remolar et al.

104 2003]. Indeed, it has been suggested that the ferric iron and sulfate mineralogy

5 105 produced by the acidic waters may be analogous to conditions experienced in Terra

106 Meridiani during the Hesperian [Fernandez-Remolar et al., 2004], and similar

107 conditions may also have occurred in portions of the Valles Marineris [e.g., Bibring

108 et al., 2007].

109

110 In this study we integrate airborne, in situ, and lab measurements to

111 characterize a suite of samples from different sites at Rio Tinto to understand the

112 mineralogical origin of the ‘doublet’ spectral signature, and we then compare our

113 findings to reflectance spectra acquired by the CRISM instrument for deposits in

114 Valles Marineris, Mars. We demonstrate that the doublet absorptions observed in

115 samples from Rio Tinto are comparable to the spectral signatures observed in Valles

116 Marineris, that the former are the result of Al-bearing phyllosilicates mixed with

117 jarosite, and that similar mineral assemblages may be present on Mars. This

118 indicates Rio Tinto contains materials that can serve as mineralogical and spectral

119 analogs for the martian deposits, though the geological processes that lead to the

120 formation of these assemblages on Earth and Mars need not be equivalent.

121 Specifically, although jarosite and Al-phyllosilicates may co-occur in geologic units

122 today on Mars, this association does not require their formation environments to

123 have been the same in space and/or time. Both mineral types may be authigenic or

124 detrital, and detailed geologic context is necessary to distinguish between these

125 options and the associated implications for local paleoenvironmental conditions.

126

127 2. Background

6 128 2.1 Overview of Clay and Sulfate Detections on Mars

129 The most commonly observed spectral signatures on Mars that have been

130 interpreted as clay minerals are identified as Fe/Mg smectite, but mixed-layer clays,

131 chlorite, Al-smectite, mica, kaolinite, and serpentine have also been observed

132 [Mustard et al., 2008; Murchie et al., 2009; Ehlmann et al., 2011; Carter et al., 2013].

133 Aqueous alteration of basaltic protoliths both at the surface and in the subsurface

134 have been invoked as the major source of these clay minerals, although subsequent

135 erosion and eolian/fluvial transport has led to their occurrence in a number of

136 different depositional environments. Intriguingly, these clay minerals, hereafter

137 referred to simply as ‘clays’, occur predominantly in terrains that are Noachian

138 (>3.6 Ga) in age [Bibring et al., 2006]. The presence and apparent spatial and

139 temporal distribution of these clays has been used to infer that early Mars was

140 characterized by circum-neutral pH, low salinity, and possibly habitable conditions

141 [Ehlmann et al., 2011]. More recently, in situ observations of clay-bearing

142 mudstones in Gale Crater by the Curiosity rover are also supportive of a habitable

143 environment, possibly during the Hesperian [Grotzinger et al., 2014; 2015; Vaniman

144 et al., 2014; Bristow et al., 2015].

145

146 In contrast, sulfate detections on Mars typically occur in younger Hesperian

147 terrains and may be indicative of a decreased hydrological cycle [Bibring et al.,

148 2006]. Decreased water availability, coupled with increasing salinity, may indicate

149 near-surface conditions inhospitable to most life [e.g., Tosca et al., 2008a]. Though

150 many sulfate salts (e.g., Ca, Na, and Mg-sulfates) can form under a wide range of pH

7 151 conditions, it is worth noting that Al3+ and Fe3+-sulfates are indicative of low-pH

152 fluids. As such, understanding the distribution and geologic setting of these minerals

153 on Mars has important implications for deciphering ancient aqueous chemistry,

154 whether it be indicative of surface waters, groundwater, or diagenetic fluids.

155

156 Jarosite, a ferric sulfate that forms under acidic conditions and that was first

157 detected on Mars by the Opportunity Rover in [Squyres et al.,

158 2004], has received considerable attention in this context. Integrated payload

159 observations by the rover reveal that jarosite is likely associated with Mg and Ca-

160 sulfates at this location, all of which occur in eolian sandstones that have been

161 interpreted as reworked dune and playa sediments [Grotzinger et al., 2005]. The

162 Curiosity rover has also detected small amounts of jarosite in clay-bearing

163 laminated mudstones in Gale Crater [Cavanagh et al., 2015], though the origin of this

164 phase (e.g., detrital or authigenic) in these deposits is not yet well constrained.

165

166 Orbital observations of jarosite and other possible ferric sulfates have also

167 been reported in the , , , and Valles

168 Marineris regions of Mars [Bibring et al., 2007; Milliken et al., 2008; Farrand et al.,

169 2009; Metz et al., 2009; Lichtenberg et al., 2010; Weitz et al., 2010; Thollot et al.,

170 2011; Noe Dobrea et al., 2011]. Of the Valles Marineris detections, possible

171 submarine fan deposits in and strata exposed on the surrounding

172 plains are notable in that ferric sulfates are found in proximity to other hydrated

173 minerals such as opaline silica, providing additional constraints on local aqueous

8 174 geochemistry [Milliken et al., 2008; Metz et al., 2009]. These rover- and orbiter-

175 based observations indicate ferric sulfates may be common alteration or diagenetic

176 phases on Mars, as discussed by Burns [1988; 1993], and understanding their

177 occurrence can provide insight into local, regional, or global redox conditions.

178

179 2.2 CRISM Observations of the Spectral ‘Doublet’

180

181 CRISM spectra of martian deposits that exhibit the ‘doublet’ feature

182 mentioned above are characterized by absorption maxima (reflectance minima)

183 occurring at wavelengths of 2.205-2.218 and 2.265-2.278 μm. These spectral

184 features have been noted previously in a number of CRISM-based studies,

185 particularly for deposits within Valles Marineris, and in some of these studies the

186 material has been referred to as a hydrated silicate [e.g., Roach et al., 2010].

187 However, many minerals, including gypsum, jarosite, nontronite, and silica, exhibit

188 absorptions in this wavelength range, and it is possible that the spectral doublet

189 represents a mixture of two or more hydrated minerals (Figure 1). The

190 doublet feature is consistent with vibrational absorptions due to the presence of

191 cation-OH bonds, although the position and shape of absorptions within the feature

192 lead to ambiguity in determining the specific cation(s). The shorter wavelength

193 absorption (~2.21 µm) may be consistent with Si- or Al-OH bonds, whereas the

194 longer wavelength feature (~2.27 µm) may be indicative of Fe- or Mg-OH bonds.

195 However, spectra of H2O-bearing Ca-sulfates such as gypsum also exhibit

9 196 absorptions in this wavelength region, despite the lack of structural Si, Al, Fe, Mg, or

197 OH.

198

199 In addition, CRISM spectra that exhibit the doublet feature also exhibit an

200 absorption at ~1.9 µm due to the combination stretch+bend vibration mode of H2O.

201 A similar doublet absorption has also been observed in laboratory spectra of poorly

202 crystalline Fe-silicate and sulfate mixtures [Tosca et al., 2008b] as well as acid-

203 leached Fe-smectite [Madejova et al., 2009], suggesting the spectral signature may

204 also be consistent with poorly crystalline or amorphous phases. Therefore, while it

205 is likely that the spectral doublet feature is indicative of OH and/or H2O-bearing

206 phases, the specific mineral assignments remain ambiguous.

207

208 Roach et al. [2010] proposed four hypotheses to explain the doublet

209 material: (1) a mixture of sulfates (in particular, sulfates structurally similar to

210 gypsum and jarosite), (2) a mixture of opaline silica and Fe/Mg-smectite, (3) an Al-

211 OH phase mixed with Fe/Mg-smectite and/or jarosite, and (4) acid-leached Fe/Mg-

212 smectite or neoformation of poorly crystalline clays. Subsequent study by Weitz et

213 al. [2014] supported the latter two hypotheses based on an analysis of CRISM

214 spectra for the Melas Chasma region, demonstrating that the spectral doublet is

215 associated with “light-toned draping deposits” (termed LD in Weitz et al., 2014) that

216 locally form some of the highest stratigraphic units. These LD deposits extend

217 through portions of Melas Chasma and drape over pre-existing topography,

218 suggesting an airfall origin. Weitz et al. [2014] interpreted this LD unit to have been

10 219 deposited after the Late Hesperian valleys that incise the walls of the chasma, but

220 stratigraphic position alone cannot be used to determine whether the hydrated

221 phase(s) associated with the spectral doublet in this unit represent transported

222 sediments or authigenic products.

223

224 Roach et al. [2010] described a similar thin light-toned unit in Ius Chasma

225 that drapes pre-existing topography, but the stratigraphic and possible genetic

226 relationships between this unit and those in Melas have not yet been established.

227 Roach et al. [2010] interpreted the light-toned unit to have been emplaced on top of

228 Fe/Mg-smectite and sulfate-bearing strata either during or after a period of acidic

229 aqueous alteration when sulfates formed from evaporating brines. Another study by

230 Weitz et al. [2011], of Noctis Labrynthis, identified the doublet material in a ~20 m-

231 thick massive light-toned unit that outcrops near the top of a stratified light-toned

232 deposit [see Weitz et al., 2011 Figure 2]. It was suggested that this deposit was also

233 formed under acidic conditions, and the doublet-bearing unit was interpreted to

234 pre-date deposition of Fe/Mg- smectite and sulfates in the same region [Weitz et al.,

235 2011]. Although the doublet-bearing materials in Valles Marineris are often

236 associated with units that drape pre-existing topography, including the canyon

237 wallrock, Metz et al. [2010] provided evidence for deformation features in some of

238 the deposits, suggesting they may have originated higher up the canyon walls than

239 indicated by their current topographic position. Finally, CRISM data for Mawrth

240 Vallis also show evidence for doublet-bearing materials in the vicinity of Fe/Mg-

241 smectites and Al-phyllosilicates [Noe Dobrea et al., 2011; Bishop et al., 2013].

11 242

243 To summarize, doublet-bearing units are present in stratigraphic sections

244 that contain hydrated minerals suggestive of complicated and extended periods of

245 water-rock interaction and/or sediment transport. Fe/Mg-smectites and sulfates are

246 often found in nearby units and in most cases these occur in strata emplaced prior

247 to the doublet-bearing units. An exception is Noctis Labrynthis, where the doublet

248 material occurs in a unit that lies stratigraphically below an Fe/Mg smectite-bearing

249 outcrop.

250

251 We note that jarosite, which typically forms under conditions of pH<4, is also

252 identified in several of the regions that also exhibit the spectral doublet [Milliken et

253 al., 2008, Metz et al., 2009, Noe Dobrea et al., 2011]. Although the jarosite deposits

254 are not necessarily in rocks that are stratigraphically adjacent to the doublet

255 materials, the presence of this mineral in these regions may provide insight into

256 local aqueous geochemistry. Jarosite deposits on Mars may have formed where

257 acidic sulfate-bearing water ponded and subsequently evaporated [Tosca and

258 McLennan, 2006], though it is also possible for jarosite to form through chemical

259 weathering of [e.g., Madden et al., 2004]. Alternatively, acidic conditions that

260 lead to jarosite formation may be induced by local or regional oxidation of Fe2+-

261 bearing fluids, possibly via the emergence of groundwater and interaction with the

262 oxidizing martian atmosphere [Hurowitz et al., 2009]. An important aspect of any of

263 these scenarios is that the presence of authigenic jarosite records information about

264 local pH as well as redox conditions.

12 265

266

267 2.3 Overview of Field Region at Rio Tinto, Spain

268 The headwaters of the Rio Tinto near Nerva, Spain are acidic (mean pH of

269 2.3) and rich in ferric iron and other heavy metals. Three terraces of the Rio Tinto

270 are exposed in the river valley with the youngest terrace closest in elevation to the

271 modern day channel and the intermediate and oldest terraces located

272 stratigraphically above it (see Figure 2). The youngest of these (called the lower or

273 young terrace) is Holocene in age (750-1500 years) and composed of poorly sorted

274 conglomerates overlain by cross-bedded sandstones. Both types of deposits are

275 consistent with formation by mass wasting and sediment accumulation in the active

276 channel, a process that also occurs in the modern river. The Pleistocene-aged (~35

277 Ky) intermediate terrace is exposed 100 m above the lower terrace and consists of

278 poorly sorted clasts of country rock within a goethite-bearing matrix; this terrace is

279 distinct from the other river terraces as it lacks the sedimentary facies

280 corresponding to channel and backwater deposits seen in both the young and old

281 terraces. Finally, the Late Pliocene-aged (2 My) upper or old terrace is exposed in a

282 ≥10m thick outcrop at Alto de la Mesa. The outcrop at Alto de la Mesa consists of a

283 lag conglomerate overlain by sandstones and iron oxide precipitates, reflecting

284 similar types of channel deposits as those seen in the lower terrace. Further

285 description of the terraces and radiometric age constraints are reported in

286 Fernández-Remolar and Knoll [2008].

287

13 288 The mineralogy of the modern river system and its terraces is dominated by

289 ferric oxides and ferric sulfates. A number of previous studies have investigated

290 variations in the mineralogical assemblages in the river terraces using X-ray

291 diffraction (XRD), visible-near infrared (VIS-NIR) reflectance spectroscopy, and

292 Raman spectroscopy [Fernández-Remolar et al., 2005; Roach et al., 2006; Chemtob

293 et al., 2006; Sobron et al., 2009, 2014,]. The reader is directed to Fernández-Remolar

294 et al. [2005] for a more thorough discussion on the mineralogy and mineral-forming

295 processes of the Rio Tinto. Those authors found that seasonal precipitation of Ca-

296 and Fe-sulfates occurs near the source region of the river, with the Fe-sulfates

297 jarosite (in particular hydronium jarosite), schwertmannite, and copiapite making

298 up much of the mineral assemblage, along with nanophase iron oxide and

299 nanophase goethite. Only the most stable minerals (iron oxides, goethite, jarosite)

300 survive through the wet season, during which meteoric water dissolves many of the

301 precipitates and aids in cementation of the terrace deposits. Jarosite is less

302 abundant and often absent in the older deposits, likely due to its long-term

303 instability under wet conditions, and these units instead contain more stable Fe-

304 oxides such as goethite and hematite. Therefore, the persistence of jarosite may

305 indicate a low degree of water-rock interaction or low water-to-rock ratios since its

306 precipitation. Country rock in this region is composed of quartz, chlorite, and pyrite,

307 which act as a source of solutes for the precipitates as well as a source of detritus for

308 the terrace conglomerates.

309

14 310 In addition to sulfates and Fe-oxides, phyllosilicates (2M1 illite, chlorite,

311 kaolinite, and smectites) are also present throughout the Rio Tinto system and are

312 thought to form by alteration of the volcanosedimentary bedrock near the source of

313 the acidic river system [see Fernández-Remolar et al., 2011 for details]. At Rio Tinto,

314 phyllosilicates typically accumulate at times of maximum waterflow during the wet

315 season when deposition of detrital sediment is dominant over chemical

316 sedimentation, but they can also be found in environments where conditions are

317 outside those typical for clay formation, such as the low-pH source streams. The

318 diversity and stability of these phyllosilicates depends on the hydrology, source rock

319 composition and geochemical processes occurring in the river, all of which must be

320 accounted for when interpreting the deposits [Fernández-Remolar et al., 2011].

321 Likewise, jarosite is neither stable nor formed in all of Rio Tinto’s hydrogeochemical

322 environments [Fernández-Remolar et al., 2005 and discussed below].

323

324 3. Methods

325 3.1 Hyperspectral Imaging Data

326 3.1.1 HyMap Data Analysis

327

328 For remote observations of Rio Tinto we used aerial HyMap scenes that

329 covered the headwaters of the river, nearby mining activity, and urban areas. These

330 data were acquired by Integrated Spectronics on Aug 14, 2004 and are the same as

331 the fully processed and analyzed data used by Roach et al. [2006]. The reader is

332 directed to that publication for a detailed description of how these data were

15 333 calibrated, atmospherically corrected, and processed. Briefly, HyMap is an aerial,

334 hyperspectral system with 126 channels covering a wavelength range of 0.45 to 2.5

335 μm at a spatial resolution of ~8m per pixel. Spectra from the processed images were

336 examined for individual pixels as well as spectral averages from multiple pixels that

337 correspond to larger regions of geologic or spectral interest, called regions of

338 interest (ROIs). Wavelengths from ~1.3 - 1.45 μm and ~1.8 - 2 μm were excluded

339 due to atmospheric absorptions. A principal component analysis (PCA) was carried

340 out for the data to reduce spectral variance and to highlight regions with similar

341 spectral properties (Figure 3). PCA is a purely mathematical procedure and does not

342 rely on a user’s a priori knowledge of variability in a region; thus, it can be difficult

343 to attribute different features observed in PCA images to unique differences in

344 mineralogy. However, as a purely qualitative tool we found this method reliable for

345 highlighting regions in the HyMap scenes whose spectra exhibited the doublet

346 absorption feature (e.g., Figure 3).

347

348 3.1.2 CRISM Data Analysis

349

350 CRISM is a hyperspectral imaging system on NASA’s Mars Reconnaissance

351 Orbiter that consists of several spectrometers. The “L” spectrometer covers near-

352 infrared wavelengths from 1.0 to 3.9 μm using 431 channels, has a spectral

353 resolution of ~6.55nm per channel, and a maximum spatial resolution of ~15-19 m

354 per pixel [Murchie et al., 2007]. The highest spatial-spectral resolution images are

355 Full Resolution Targeted (FRT) observations, and for this study we analyzed three of

16 356 these images in detail from the western Valles Marineris region near Melas and Ius

357 Chasma (see Figure 4, Figure 9). A full description of CRISM image processing

358 methods is provided in the supplemental material.

359

360 The variability in the doublet absorption feature was studied in detail by

361 handpicking spectra from geologic units and ROIs that exhibited these spectral

362 features. The resulting spectra commonly represent 3x3 or 5x5 pixel averages.

363 Each spectrum was fit with an upper convex hull to define the spectral continuum

364 over the ~2.1 to 2.35 μm wavelength range. The original spectra were then divided

365 by these continuum fits to produce continuum-removed spectra. This step isolates

366 the wavelength region where the doublet occurs and is necessary to compare band

367 depths between spectra [ and Roush, 1984]. Band depth is defined as 1-[Rb(λ

368 )/Rc(λ)], where Rb represents the observed reflectance value at wavelength λ and

369 Rc represents the calculated reflectance of the spectral continuum fit at the same

370 wavelength. Band depth maps were created for each CRISM image to provide a more

371 in-depth analysis of the spatial distribution of the spectral doublet feature.

372

373 Gaussian curves were also fit to the continuum-removed doublet absorptions

374 and the resulting Gaussian amplitudes and center positions were mapped for each

375 CRISM image. Gaussian modeling of absorptions is commonly used for remote

376 sensing data to assess band strength in the presence of noise or to remove the

377 influence of adjacent or partially overlapping absorption features on the absorption

378 of interest [Clark and Roush, 1984; Sunshine and Pieters, 1993]. Noise in the CRISM

17 379 images, for instance, may introduce apparent changes in band position/strength if

380 these calculations are based on the observed absolute reflectance values on a pixel-

381 by-pixel basis. This can introduce uncertainties in the band depth maps, which we

382 attempt to mitigate by examining the amplitudes and positions of Gaussian fits.

383

384 To determine the background noise in an image prior to Gaussian fitting, a

385 histogram of the standard deviations within each image was examined for a flat,

386 absorption-free portion of the spectrum (1.70-1.85 μm), an example of which is

387 presented in Figure 4. Here it is clear that the standard deviations in the ratio

388 spectra are predominantly <0.005, thus any absorption feature (band depth) in the

389 ratio spectra that is larger than this value is above the typical level of ‘noise’ in the

390 CRISM image. Two CRISM ratio spectra exhibiting the doublet feature and their

391 corresponding Gaussian fits are also presented in Figure 4, where it is clear that the

392 doublet absorptions are much stronger than the typical channel-to-channel

393 variation in reflectance. Pixels that exceed the 0.005 threshold are fit with two

394 Gaussian curves, one for each absorption in the doublet (see example in Figure 4).

395

396 3.2 Field Measurements

397 Fieldwork was conducted at multiple locations in the vicinity of Rio Tinto,

398 Spain in early May of 2014. Rock and/or soil specimens were collected and field

399 spectra were acquired at each of five primary sites (Table 1). These locations

400 included the modern streambed environment and two successively older terraces

401 that have been affected to different degrees by diagenetic processes (Figure 2).

18 402 Previous studies have shown that the materials at these sites exhibit different

403 mineral assemblages [Fernández-Remolar et al., 2005; Fernández-Remolar and

404 Knoll, 2008], allowing us to examine whether spectral reflectance properties can be

405 linked to post-depositional processes in these types of environments.

406

407 Samples collected were collected from the Rio Tinto source region (Anabel’s

408 Garden, Richi Spring, and Source Spring sites), a series of small streams with

409 actively forming efflorescent and crust-forming precipitates (e.g. copiapite, jarosite,

410 gypsum). A number of samples were also collected from an actively mined terrace

411 directly above Anabel’s Garden that showed evidence for the spectral doublet

412 absorption in the airborne HyMap data. Another field site, Berrocal, is also located

413 directly on the Rio Tinto approximately 20 km downriver from the mining

414 operations. At this location, there is a mix of detrital sediment and authigenic

415 precipitates. We also examined older deposits at Barranco de los Locos, a site that is

416 part of the Holocene-aged lower terrace in which the lower unit is a ferruginous

417 conglomerate and is overlain by sandstone [Fernández-Remolar et al., 2005].

418 Finally, for comparison to the modern environment we examined an older (early

419 Quaternary) terrace exposed at Alto de la Mesa. Detailed descriptions of these sites,

420 their ages, and local geology can be found in previous publications [e.g., Fernández-

421 Remolar et al., 2005; Fernández-Remolar and Knoll, 2008; and references therein].

422

423 A portable VIS-NIR spectrometer (ASD FieldSpec 3®) was used to acquire

424 reflectance spectra in situ from 0.35 to 2.5 μm. Reflectance spectra were acquired

19 425 for a variety of materials passively (relying on ambient sunlight) and/or actively

426 (relying on a contact probe with internal illumination). All measurements were

427 made using a Spectralon® panel as the reflectance standard. Passive measurements

428 that rely on solar illumination and reflectance are more directly analogous to

429 airborne data such as HyMap. A more detailed description of the ASD measurements

430 is provided in the supplemental material.

431

432 Rock and sediment samples were collected to capture the spectral diversity

433 of deposits at each location and were based in part on interesting absorption

434 features observed in the HyMap and field reflectance spectra (see Table 1 for full

435 list). To the extent possible, samples were collected at the different sites to be

436 representative of the mineralogy expected at HyMap and CRISM spatial scales (e.g.,

437 the dominant mineral assemblages at scales of ~8 - 18 m/pixel, respectively),

438 focusing on regions whose spectra exhibited evidence for compositional diversity.

439 Of specific interest were regions with spectral or field-based evidence for Fe-oxides,

440 Fe-sulfates, and clay minerals. A number of locations contained materials whose

441 reflectance spectra exhibited the doublet absorption feature described above, and a

442 variety of these materials were collected for detailed laboratory characterization.

443 Though many other samples and spectra were collected during the field campaign,

444 the samples, field measurements, and HyMap data whose spectra exhibit these

445 distinct doublet absorptions were chosen as the focus for the present study. A more

446 complete treatment of how field-based spectral observations and diversity at Rio

20 447 Tinto relate to aerial and orbital-scale observations is presented in Roach et al.

448 [2006].

449

450 3.3 Laboratory Measurements

451 3.3.1 Sample Preparation

452 All laboratory analyses were conducted in the Department of Earth,

453 Environmental and Planetary Sciences at Brown University. A number of small chips

454 (~2.5 cm) were removed from each hand sample and a second portion of each was

455 ground by hand in an agate mortar and pestle. The resulting powders were ground

456 such that the material could pass through a <45 μm mesh sieve. Grinding and

457 sieving serves the dual purpose of homogenizing the sample and reducing particle

458 size-dependent absorption/scattering processes. Where possible, separate powders

459 were prepared for different portions of a single sample to assess heterogeneity (e.g.

460 rock varnish, dark or light components, areas with different degrees of weathering)

461 and for comparison to the whole rock powders. Some hydrated mineral phases are

462 expected to be unstable even over short time frames, and these are not included in

463 our dataset.

464

465 3.3.2 Reflectance Measurements

466

467 Reflectance spectra of the hand samples, chips, and powders were measured

468 in the laboratory using the same ASD FieldSpec 3 instrument over a wavelength

469 range of 0.35 to 2.5 μm. The measurement spot size was adjusted to encompass the

21 470 entire sample in order to provide a better assessment of the bulk spectral properties

471 of each hand, chip, and powder sample rather than averaging spectra of individual

472 components. We note that the chips and powder samples were placed in ~10 mm

473 diameter sample holders for measurement, thus the measurement spot size for

474 these samples is equivalent.

475

476 All of the rock powders sieved to <45 µm were also measured with a Nicolet

477 iS50 FTIR from 1.75 – 25 μm. The FTIR was equipped with a Praying Mantis diffuse

478 reflectance attachment from Harrick Scientific and diffuse from Labsphere was

479 used as a reflectance standard. These measurements used a KBr beamsplitter and

480 DTGS detector, and each final spectrum was an average of 50 scans. Absolute

481 reflectance values from the FTIR are highly dependent on any slight differences in

482 height between the sample and gold standard, thus all FTIR spectra were scaled to

483 match the ASD values at 1.7μm. These spliced spectra provide full VIS-NIR-mid-IR

484 wavelength coverage, and because all samples were measured under identical

485 conditions, the observed relative differences between samples are reliable.

486

487 All reflectance spectra (field and lab) were processed using the same

488 methods, and continuum-removed spectra and band depths were calculated using

489 the same procedures as described above for the CRISM data (i.e., an upper convex

490 hull method was used to define the continuum slope). We analyzed the continuum-

491 removed spectra using Matlab to find the local minima (which is equivalent to the

492 maximum band depth) within the ~2.1 – 2.35 µm wavelength region. If the

22 493 continuum-removed spectrum is determined to exhibit absorptions near ~2.2 μm

494 and ~2.26 μm, then this so-called doublet feature is further analyzed to determine

495 the total width, center position, and depth of the individual absorptions.

496

497 3.2.3 X-ray Diffraction Measurements

498

499 The <45μm sieved powders were measured with a Bruker D2 Phaser XRD

500 system (see supplemental material for details) to determine qualitative and

501 quantitative mineralogy. Samples whose corresponding reflectance spectra

502 exhibited the 2.2 μm doublet (n = 8) were studied in detail, as were a subset of

503 samples from multiple locations whose spectra did not exhibit the doublet (n = 3) in

504 order to test our hypothesis that a distinct mineral assemblage gives rise to the

505 doublet feature (Table 2). Powder diffraction patterns were typically acquired over

506 a range of 2-90° 2θ, 0.02˚ step size, with a total integration time of up to ~12 hours

507 (6.5 s/step).

508

509 XRD patterns were analyzed with two programs, DIFFRAC.SUITE™ EVA for

510 full pattern phase identification and DIFFRAC.SUITE™ TOPAS for quantitative

511 Rietveld refinement, where the latter is similar to the method described in Bish and

512 Howard (1988). An overview of the mineralogy at Rio Tinto is provided in

513 Fernández-Remolar et al. [2005], which we used as a guideline for choosing

514 reasonable starting components for the qualitative mineral assessment. Modal

515 mineralogy estimated from Rietveld refinement relied on the structural and

23 516 instrumental parameters to model the XRD patterns. Structure files describing the

517 crystallographic structure of relevant minerals used in the refinement are listed in

518 the supplementary table.

519

520 These mineral structures were added and removed from the model

521 iteratively until a reasonable fit to the observed pattern was obtained. TOPAS

522 automatically normalizes the model results such that mineral abundances will

523 always add to 100%. The quality of the refinement is determined by the weighted

524 profile R-factor (Rwp), which considers the root mean square error (RMSE) for each

525 scaled intensity value (peak and background). Although a smaller Rwp generally

526 indicates a better model, it is possible to over fit the data with this method and

527 discretion is used to ensure the resulting model is geologically realistic. We also

528 calculated the areas for the muscovite 001 (~9° 2θ) and jarosite 101 (~15° 2θ)

529 peaks. Though this method does not provide quantitative estimates like the Rietveld

530 method, it does provide an independent estimate of how the relative abundances of

531 the two minerals vary between samples without the uncertainties associated with

532 modeling the full XRD pattern.

533

534 4. Results

535 4.1 Overview of Rio Tinto Field Observations

536 Previous work has shown that mineralogical and spectral differences exist

537 between the old and young terraces [Fernández-Remolar et al., 2005; Sobron et al.,

538 2014], and we observe congruous spectral differences in our samples from these

24 539 locations. Field spectra of samples from the source region exhibit absorptions

540 consistent with the presence of gypsum, jarosite, and copiapite (Figure 5). Hydrated

541 iron-bearing phases are also common in these field spectra, including absorptions

542 consistent with the presence of schwertmannite and copiapite, which are identified

543 in spectra from the Richi Spring and Source Spring locations. The field spectra from

544 Berrocal (not shown) have similar identifying absorptions that are typical of

545 the precipitates, phyllosilicates and iron hydroxides found in the source region

546 spectra. Spectra exhibiting the doublet absorption are found in sand bar sediments

547 from Berrocal, representing rocks/sediments that have not been influenced by

548 mining activities.

549

550 Most of the Fe- and Ca-sulfate precipitates are absent at the young terrace

551 site, Barranco de los Locos. Jarosite, Al-phyllosilicate (muscovite/illite), and goethite

552 are the primary minerals identified in our field spectra at this location. VIS-NIR

553 spectra from the old terrace site, Alto de la Mesa, exhibit the lowest mineral

554 diversity and are dominated by iron oxides. Most spectra from Alto de la Mesa are

555 consistent with the presence of goethite, though some indicate the presence of

556 kaolinite or chlorite. Sulfate absorptions are not apparent in spectra of materials

557 from the old terrace, and the doublet feature is also not observed in these spectra.

558 However, spectra for a number of samples that were taken from a mine road terrace

559 above Anabel’s garden display the doublet absorption. Other samples from this site

560 are spectrally consistent with goethite and hematite. Finally, samples from some of

25 561 the mining terraces are spectrally consistent with chlorite, kaolinite, and muscovite,

562 likely reflecting country rock components.

563

564 4.2 Laboratory Measurements

565 4.2.1 X-ray Diffraction

566

567 Based on an initial qualitative assessment of the XRD patterns and observed

568 spectral diversity, eleven samples were chosen for further quantitative Rietveld

569 refinement. Eight of the samples have spectra that exhibit the spectral doublet and

570 contain between 3 - 38% jarosite and 4 - 50% illite or muscovite (Table 2). The

571 other volumetrically important mineral in these samples is quartz, which does not

572 have absorptions at VIS-NIR wavelengths, so the samples are spectrally dominated

573 by Fe-sulfate and Al-bearing clay minerals. K-jarosite is a better fit in the Rietveld

574 refinement than either hydroniumjarosite or natrojarosite. Though the mineral

575 assemblage is largely the same, the relative proportion of jarosite and Al-clay is

576 quite variable between samples (Figure 6), and these values are presented in Table

577 2. Three samples that do not show evidence for the doublet feature in

578 corresponding reflectance spectra were also examined. The XRD patterns for these

579 samples are consistent with a mineral assemblage dominated by hematite, goethite,

580 kaolinite, chlorite and quartz, but no jarosite, in agreement with the findings of

581 Fernández-Remolar et al., [2005] for this locale (Table 2). One sample from the

582 young terrace, Barranco de los Locos sample RT23, is modeled to contain ~33%

583 jarosite and ~39% gypsum but no Al-clay (Table 2), and its reflectance spectrum is

26 584 dominated by gypsum at wavelengths >1.3 µm (Figure 5). In summary, the major

585 secondary phases in these samples are jarosite, illite or muscovite, and Fe-oxides

586 such as goethite and hematite. Although smectitic clays are an integral part of the

587 alteration mineral assemblages on Mars, the Rio Tinto samples examined for this

588 study show no clear evidence for the presence of smectitic clays.

589

590 4.2.2 Reflectance Spectra

591

592 As with the field spectra, the laboratory spectra indicate variations in

593 mineral assemblage with age. In samples from the modern streambeds (samples

594 from Anabel’s Garden and Source Spring) gypsum is the primary mineral identified

595 in two samples (RT18, RT23), with jarosite also being present in RT23 (see Figure

596 5). Other samples from these sites have spectra with strong OH and H2O bands at

597 ~1.4 and ~1.9 µm as well as absorptions consistent with copiapite, schwertmannite,

598 or possibly monohydrated sulfates (Figure 5). These hydrated samples were not

599 measured with XRD because these minerals change hydration states readily under

600 laboratory conditions, so these spectral mineral identifications cannot be confirmed.

601 However, other XRD studies of Rio Tinto samples have found evidence for a variety

602 of sulfates, including epsomite, copiapite, and coquimbite [Fernández-Remolar et al.,

603 2005], and it is likely that a range of hydrated sulfates are also present in our

604 samples.

605

27 606 The spectral shapes of samples from the young terrace are distinct from

607 those of Anabel’s Garden and Source Spring, with the former having weaker

608 hydration absorptions and a strong spectral slope near 1 μm. In fact, the major

609 difference in spectra of all the young terrace samples lie in the region between 2.2

610 and 2.3 μm; spectra of some samples show a doublet in this region whereas others

611 have only an ~2.21 or 2.27 μm absorption. Samples collected from the old terrace

612 deposit are spectrally consistent with clays such as kaolinite and chlorite as well as

613 Fe-oxides such as hematite (Figure 5). Gossan samples all show spectral evidence of

614 goethite. As with the field spectra, lab measurements of the older samples typically

615 lack spectral evidence for sulfates and are instead indicative of Fe-oxides and/or

616 phyllosilicates. Samples from the mining terrace have spectra that exhibit the

617 spectral doublet or that are representative of the same types of minerals discussed

618 for the other sites, including chlorite and Fe-oxides (goethite and hematite).

619 However, given that the specific geologic context of these samples is unknown,

620 especially compared to the well-defined mineralogical and age progression of the

621 river terraces, it is difficult to comment on the processes that have affected this

622 group of samples.

623

624 To summarize thus far, laboratory spectra with the doublet absorption

625 correspond to samples collected from the modern stream sites (Anabel’s Garden,

626 Source Spring), the young terrace (Barranco de los Locos), and from a mining

627 terrace site near Anabel’s Garden (Table 2). There are many minerals that have

628 spectral features near 2.26 µm, but we attribute the 2.26 µm feature in the doublet

28 629 to jarosite because of the combined presence of absorptions at wavelength positions

630 that are attributed to jarosite (1.48, 1.86, 1.9, 2.26 µm), as well as evidence from

631 XRD mineralogical analysis. The doublet feature is conspicuously absent in spectra

632 of samples collected at the old terrace, all of which lack jarosite. Reflectance spectra

633 of the <45 µm powders of samples exhibiting the doublet show strong, broad

634 electronic transition absorptions at wavelengths of ~1 μm due to the presence of

635 iron (Figure 6). There are also narrower vibrational absorption features at 1.4 and

636 1.48 μm due to OH, though the 1.48 μm feature is weak or absent in samples that

637 show a very strong 2.21 μm short wavelength band in the doublet feature. Likewise,

638 a broader absorption at ~1.9-1.95 μm due to OH/H2O and a narrower absorption at

639 1.86 μm from OH/H2O are also weaker in samples with a stronger ~2.21 µm

640 feature.. In contrast, the ~2.21 µm absorption in the doublet feature is related to the

641 presence of Al-phyllosilicates (illite or muscovite). A vibrational absorption near

642 2.35 μm is present in some samples and is also consistent with the presence of Al-

643 OH in illite or muscovite.

644

645 Spectra of rock chips and hand samples were also acquired and, in most

646 cases, the samples whose powder spectra exhibited the doublet feature also exhibit

647 the feature for these larger physical forms (see example for sample RT19 in Figure

648 7). However, sample RT10 is an exception in that a spectrum of the orange-colored

649 varnished face of the hand sample shows the doublet feature whereas the

650 unvarnished side of the hand sample, a chip, and the powder spectra are dominated

651 by the 2.21 μm feature (Figure 7). This indicates the varnish may be enriched in

29 652 jarosite, and although such coatings or varnishes may be only several micrometers

653 thick, they may contribute significantly to aerial and orbital spectral observations.

654

655 Only the bulk powder versions of the samples were measured using both

656 reflectance spectroscopy and XRD; thus, we compare only the powder spectra

657 directly with the mineralogy determined by XRD. There is a strong correlation

658 between the strength of the longer wavelength absorption in the doublet feature

659 and the abundance of jarosite for the samples examined here (Figure 8). There is

660 also a strong relationship between the relative band strengths and the relative

661 amounts of jarosite and Al-clay in the samples, as determined using either the

662 Rietveld modal estimates or the XRD peak area. The relationship between Al-clay

663 and the strength of the shorter wavelength absorption has a lower R2 value than in

664 the other plots, but we suggest this is due to the size of the dataset, where a

665 relatively small amount of scatter (in this case one data point with anomalously high

666 estimated Al-clay wt.%) has a strong control on the fit. These observations indicate

667 that the doublet feature observed in spectra of some samples from Rio Tinto is

668 related to the mixing of Al-phyllosilicate (illite or muscovite) with jarosite, and that

669 the strengths of the absorptions within the doublet feature are directly related to

670 the relative proportion of these phases.

671

672 4.3 HyMap and CRISM Observations

673 4.3.1 HyMap Results

674

30 675 The HyMap PCA false-color map presented in Figure 3 highlights pixels

676 whose spectra exhibit the doublet feature in dark red. Absorption maxima

677 (reflectance minima) for these pixels occur at ~2.20-2.22 and 2.26-2.28, a range that

678 is broader than observed in Mars CRISM data and that could be a result of

679 differences in spectral resolution, since HyMap only has 6 bands in this wavelength

680 region compared to CRISM’s 13 bands. The doublet feature is evident in HyMap

681 spectra corresponding to the Anabel’s Garden location, where it is possible to track

682 changes in the relative strength of the absorptions across a ~500 m transect in the

683 HyMap image (Figure 3). The feature is also present in numerous pixels

684 corresponding to mining terraces, although it was not possible to independently

685 examine most of these locations during the field study. Bedrock, detritus, and soil at

686 the young, intermediate, and old terrace sites discussed above are not always

687 apparent in the HyMap data due to the presence of vegetation. However, chlorite,

688 kaolinite, gypsum, and Al-clay are identifiable in the HyMap spectra, and Roach et al.

689 [2006] provide a thorough analysis of iron oxides and hydrated or hydroxylated

690 iron-sulfates in these airborne data. In this study we have only relied on the HyMap

691 data to provide an airborne comparison (at lower spatial and spectral resolution) of

692 the doublet-bearing materials with the spectra acquired in the lab and in situ using

693 the ASD FieldSpec3. In this context, the HyMap data confirm that spectra of certain

694 surface materials exhibit the doublet feature even at a meters-per-pixel spatial scale

695 and lower spectral resolution, consistent with the higher spatial and spectral

696 resolution field and laboratory data.

697

31 698 4.3.2 CRISM Results

699

700 The spectral doublet feature is evident in a number of CRISM images in the

701 Valles Marineris, with particularly good examples in Melas and Ius Chasma [e.g.

702 Roach et al., 2010, Weitz et al., 2014]. Our Gaussian fit analysis of three of these

703 CRISM images show that the absorption band centers and relative strengths

704 (Gaussian amplitudes) associated with the doublet materials are only moderately

705 variable (Figure 9). The centers of the Gaussian fits for the three CRISM images and

706 the HyMap image are listed in Table 3. The typical 2σ deviation on the band

707 positions in the CRISM images is ~0.015 µm and the band centers fall near 2.2060

708 and 2.2682 µm on average. However, the wavelength position varies across the

709 spatial dimension of the CRISM detector (spectral smile) by ~0.010 µm, which may

710 explain much of the width of the histograms in Figure 9. To examine this effect we

711 analyzed a 50 x 50 pixel region restricted to the center of each image where the

712 spectral smile is minimal. Histograms for these subsets reveal a decrease in

713 standard deviation as well as a shift in mean band center in all three cases (Table 3,

714 Figure 4). The difference in band centers between the full image and the central

715 subset is statistically significant (p < 0.05) in every case.

716

717 Therefore, we conclude that the band centers of the doublet spectral feature

718 are relatively constant and occur at wavelengths of ~2.21 and ~2.265-2.270 µm. It

719 should be noted that Weitz et al. [2014] identified another spectral doublet

720 occurrence with a long wavelength absorption closer to ~2.28 µm, and the inclusion

32 721 of pixels with this feature in our analysis would create additional spread in the

722 histograms in Figure 9c. Given that we did not analyze each pixel in each image

723 individually, it is possible that this longer wavelength doublet is present, but it is

724 certainly not a dominant spectral signature, which is evident in the histograms and

725 investigation of spectra from individual pixels and ROIs.

726

727 While the wavelength positions of the two absorptions appear to be fixed,

728 their relative strengths are variable and are not directly correlated (Figure 9d).

729 Though this result is qualitatively similar to what is observed in field, lab and

730 HyMap spectra of materials from Rio Tinto, the range in relative amplitudes of the

731 two absorptions is typically much smaller in the CRISM data than for the terrestrial

732 examples. That is, spectra of some materials from Rio Tinto are clearly dominated

733 by the ~2.21 or the ~2.27 µm feature, with the other feature being very weak or

734 absent. In contrast, nearly all CRISM spectra for the regions examined here exhibit

735 both features and in many instances both features are of equal strength and, while

736 variability does exist, it is less intense than the variability in strengths observed in

737 the Rio Tinto spectra.

738

739 The relative band strengths, which is the Gaussian amplitude of the 2.21 µm

740 absorption divided by the amplitude of the 2.27 µm absorption, for the three CRISM

741 images is 1.1925±1.47 (2σ), 1.2714±1.35 (2σ), and 1.2207±1.28 (2σ). These

742 averages suggest that the band strength of the shorter wavelength absorption is

743 often slightly stronger than the longer wavelength absorption. However, the

33 744 standard deviation of the distribution is large enough that it is clear the two

745 absorptions do not co-vary. That is, the two absorptions that comprise the doublet

746 feature in CRISM spectra are independent of one another and thus represent two or

747 more distinct phases. Both tails of the distributions in Figure 9d are affected by

748 “bad” pixel values (typically a spike in the value for a single CRISM channel in one

749 column of the image that will influence the Gaussian fits). This is particularly

750 obvious as a bulge in the distribution as the Band 1/Band 2 amplitude ratio

751 approaches 0 (due to Band 1 amplitudes approaching 0) as well as by the long tail as

752 the Band 2 amplitude approaches 0, starting at a band amplitude ratio of ~3.

753 However, other than the extreme tails of the distributions (which represent a

754 relatively small number of pixels), these pixels play a minimal role in the overall

755 distributions of the band amplitude ratios.

756

757 4.4 A Case Study: Doublet-Bearing Materials in Ius Chasma, Mars

758

759 This study is focused on the spectral characteristics of materials found on

760 Mars and how they relate to those in Rio Tinto as seen in field, airborne, and lab

761 data, with the goal of assessing whether or not the latter may be appropriate

762 mineralogical analogs. However, to fully understand the processes and

763 environmental conditions associated with the deposits on Mars, and whether or not

764 Rio Tinto is an appropriate process analog, it will be necessary to conduct detailed

765 mapping studies that place the martian examples in their proper geologic context.

766 We conducted a small-scale examination of MRO HiRISE and CTX images of the

34 767 doublet-bearing materials in Valles Marineris to get a sense of the stratigraphic

768 position of these deposits with respect to surrounding strata.

769

770 As discussed by Roach et al. [2010], some occurrences of the materials with

771 the doublet absorption are in proximity to Fe/Mg clay-bearing strata. However, in

772 many cases the contacts between units are obscured and bedding is not apparent,

773 making it difficult to determine the relative stratigraphic position of these units.

774 Despite these complications, it is clear that the doublet absorptions are commonly

775 associated with light-toned units that drape pre-existing topography and span a

776 range in elevation (e.g., Figure 10a) or as deposits within the topographically lower

777 portions of the chasma (e.g., Figure 10b). In some locations the doublet unit

778 surrounds the clay-bearing units, whereas in others the clay-bearing unit is located

779 higher up the canyon wall. The doublet material also occurs as matrix in

780 megabreccia deposits [Roach et al., 2010], an example of which can be seen in

781 CRISM image FRT0000A396 in Figure 9. We now discuss a case study for one region

782 in Ius Chasma where several doublet-bearing units are in proximity to Mg/Fe clay-

783 bearing strata, focusing on the relative stratigraphic position of these units in order

784 to assess the relationship between the observed mineral assemblages in a more

785 detailed geologic context.

786

787 The relative strength of absorptions in the doublet feature were mapped for

788 several CRISM observations in Ius Chasma and overlain on HiRISE and CTX images

789 (Figure 11). This map reveals both the full extent of the doublet-bearing units and

35 790 allows for further discrimination of spectral properties within and between the

791 different units. Each colored pixel meets the criteria of having band depth > 0.005 at

792 both 2.205 and 2.269 µm. Where the ratio of band depths (BD 2.205/2.269 µm) is

793 small, we interpret the unit to be enriched in the ~2.27 µm phase, and a higher ratio

794 represents the opposite, in a strictly relative sense. In some locations (Figure 11c)

795 there is a full transition between these cases within a single unit. In this example,

796 the unit is heavily brecciated and lacks clear stratification. This example fits within

797 a larger regional trend where the doublet-bearing units with the strongest ~2.27

798 µm absorption are in the , and seemingly higher up the walls of the chasm,

799 which then transitions to a topographically lower region where the doublet has an

800 absorption with two bands of equal strength, and finally to the north where the

801 ~2.21 µm absorption becomes stronger in the floor units (see the red-to--to-

802 blue northerly progression in Figure 11b). In this example it is clear that the

803 doublet-bearing unit is quite heterogeneous with regard to the strength of the two

804 absorptions, suggesting the relative abundances of the phases responsible for these

805 absorptions are also highly variable within this single geologic unit.

806

807 A second example reveals changes in composition that are consistent with

808 geomorphologic changes and that conform to geologic contacts. Three doublet-

809 bearing units (“bright”, “light”, and “moderate”-toned) and a darker, stratified,

810 Mg/Fe clay-rich unit are shown in Figure 11d, with close-up HiRISE views of the

811 contacts between these units presented in Figure 12. The Mg/Fe clay unit has tilted

812 strata (dipping approximately north/northwest, see Figure 12b) and spectra for

36 813 these units do not meet the criteria for being “doublet bearing” (with the exception

814 of a few pixels that have high band ratios and therefore higher amounts of the ~2.21

815 µm component). These stratified Mg/Fe clay units are superposed by a light-toned

816 doublet-bearing unit. The significant difference in bedding geometry and the sharp

817 nature of the contact suggests there is an angular unconformity between these units,

818 with the clay strata predating the doublet-bearing unit (Figure 12b). This light-

819 toned doublet unit has a relative band strength ratio of ~1, meaning the absorptions

820 within the doublet are of equal strength (however, this does not necessarily imply

821 that there must be equal amounts of each component). Overlying this unit is a

822 second doublet-bearing unit that is moderate in tone but that exhibits similar

823 morphological characteristics as the light-toned unit (Figure 12c). At full HiRISE

824 resolution it is apparent that both the light and moderate-toned units are stratified.

825 The moderate-toned unit has a band strength ratio between 1.2 and 1.8, indicating

826 the shorter wavelength (~2.21 µm) absorption is stronger.

827

828 The third unit in this region corresponds to pixels with band ratios <0.9,

829 consistent with it being enriched in the longer wavelength (~2.27 µm) component

830 relative to the other doublet-bearing units in this location. This unit is the brightest

831 in grayscale HiRISE and CTX images and is overlain by the light-toned unit (Figure

832 12d). In this location there is no direct contact between the dark-toned Mg/Fe clay

833 strata and this bright unit, thus the relative stratigraphic order cannot be

834 determined. However, in nearby locations it appears that this bright-toned unit that

835 is characterized by a stronger ~2.27 µm band is also stratigraphically above the

37 836 Mg/Fe clay-bearing unit, though these individual outcrops are too small to identify

837 compositionally with CRISM. Therefore, the apparent increasing stratigraphic order

838 is: Mg/Fe clay strata – (angular unconformity) – bright unit – light unit – moderate

839 unit. The associated mineralogy above the unconformity appears to decrease in the

840 longer wavelength (~2.27 µm) component and increase in the shorter wavelength

841 (~2.21 µm) component upsection, at least in a relative sense.

842

843 Together, these two examples indicate that the relationships between the

844 minerals responsible for the doublet feature and the local stratigraphy are diverse.

845 In the first example there appears to be spectral variation, and thus relative changes

846 in mineral abundance, within a single geologic unit. In the second case the dominant

847 spectral variations appear to conform to stratigraphic boundaries, indicating

848 differences in relative mineral abundance as a function of geologic unit and

849 stratigraphic position.

850

851

852 5. Discussion

853 5.1 Mineralogy

854

855 The CRISM spectral doublet feature is similar to features observed in

856 airborne, field, and laboratory spectra of materials in and around Rio Tinto, Spain.

857 As discussed above, XRD analysis of select Rio Tinto samples demonstrates that this

858 spectral doublet arises from a mixture of Al-bearing phyllosilicates and jarosite. By

38 859 analogy, CRISM spectra exhibiting the doublet feature may also arise from a mixture

860 of Al-phyllosilicates and jarosite, one of the possibilities discussed by Roach et al.

861 [2010] and Weitz et al. [2014]. This interpretation is supported because (1) the two

862 absorptions that comprise the doublet feature vary independently, suggesting at

863 least two phases are present, (2) the absorption features within the martian doublet

864 examples occur at the same wavelength positions as observed in the Rio Tinto

865 spectra and are consistent with Al-OH and Fe-OH vibrations, and (3) the doublet-

866 bearing material occurs on Mars in the vicinity of other phyllosilicate and/or

867 jarosite-bearing deposits. Therefore, we conclude that Al-phyllosilicate-jarosite

868 mixtures in the Rio Tinto may serve as spectral analogs for the doublet materials

869 found in the Valles Marineris and elsewhere on Mars. However, whether or not the

870 Rio Tinto deposits may also serve as process analogs warrants further study, as

871 discussed below.

872

873 Other suggestions for the doublet feature have included opaline silica and

874 clay, mixtures of sulfates (e.g. gypsum and jarosite), and poorly crystalline Fe-SiO4-

875 bearing phases. Reflectance spectra of opaline silica do exhibit a broad absorption

876 feature from 2.2-2.26 μm [Milliken et al., 2008], but it has been noted that this

877 feature is typically too broad to match the doublet even when mixed with Fe-rich

878 clay such as nontronite [Weitz et al., 2014]. One of our Rio Tinto hand samples,

879 RT23, provides a natural example of an intimate mixture of jarosite and gypsum in

880 which there are similar proportions of both phases and no phyllosilicates (Figure 5).

881 The reflectance spectrum of this sample is clearly dominated by the H2O absorptions

39 882 characteristic of gypsum and an absorption near 2.2-2.3 μm that is wider than

883 observed for the typical doublet feature. In addition, although gypsum exhibits very

884 narrow absorption features near ~2.22 and ~2.27 µm, these are rather weak and

885 they are superposed on the much broader and stronger feature that spans the

886 ~2.08-2.3 µm region (see gypsum spectrum in Figure 1). By contrast, the

887 absorptions in the doublet feature do not appear to be superposed on any broader

888 feature and each appears to be much stronger and more clearly defined than the

889 narrow features observed at similar wavelengths in gypsum. In summary, it does

890 not seem possible to recreate the shape of the doublet feature through a

891 combination of only gypsum and jarosite, thus we consider the doublet feature to be

892 most consistent with a mixture of an Al-phyllosilicate and jarosite.

893

894 In our Rio Tinto samples the Al-phyllosilicate is dominated by illite or

895 muscovite and is likely to be detrital, sourced from the surrounding country rock. In

896 contrast, the jarosite, which in some samples occurs as a coating, is authigenic and a

897 natural product of the Fe- and S-rich, oxidizing, acidic conditions that characterize

898 the Rio Tinto system. The decrease in jarosite with progression in age suggests it is

899 not stable over geologic time in the continued presence of water, particularly for

900 circum-neutral and higher pH fluids. Indeed, we observe the doublet feature only in

901 spectra that contain jarosite, which are primarily associated with the youngest

902 deposits such as those at Anabel’s Garden and at the Source Spring. It is notable that

903 in our Rio Tinto samples the spectral doublet feature can be directly attributed to

904 relatively well-crystalline phases as there is no clear spectral or XRD evidence for

40 905 hydroxylated poorly crystalline phases that may result from the acidic leaching of

906 clay minerals. This is not to say that such processes do not occur in these locations,

907 but rather to note that the presence of acid-leached clays or hydrated poorly

908 crystalline precipitates are not required to explain the presence of the spectral

909 doublet feature in the samples examined here.

910

911 The same may not be true of the doublet-bearing materials on Mars. There,

912 acidic leaching of basalt and/or pre-existing smectitic clays may be an important

913 process for forming alteration minerals, including some sulfates such as jarosite

914 [e.g., Tosca et al., 2004; Golden et al., 2005]. As noted above, acid-leaching of clays

915 has also been invoked to explain the presence of the spectral doublet feature on

916 Mars [Tosca et al., 2008b; Roach et al., 2010; Noe Dobrea et al., 2011; Weitz et al.,

917 2014], and the occurrence of jarosite in some locations is indicative of low-pH

918 conditions, at least at a local scale. It is possible to make a poorly crystalline Fe-SiO4-

919 bearing material through the acid leaching of nontronite, and reflectance spectra of

920 such products can exhibit absorptions with band centers similar to those observed

921 in the doublet feature [Madejova et al., 2009; Roach et al., 2010]. In this scenario, the

922 relative strengths of the absorptions may vary based on the degree of acid

923 alteration. However, spectra of residues from the acid leaching of nontronite may be

924 more consistent with Si-rich, opaline-like materials, and in these cases the full width

925 of the resulting doublet feature is slightly wider than is observed in the martian

926 examples [Tosca et al., 2008b; Altheide et al., 2010].

927

41 928 As a different example, reflectance spectra of Fe/Mg smectite-rich altered

929 from Mauna Kea, Hawaii that have undergone acid leaching exhibit a doublet

930 feature that is thought to arise from the presence of jarosite and hallyosite [ et

931 al., 2012]. In this instance the spectral doublet feature is the result of two distinct

932 (presumably authigenic) mineral phases likely produced by

933 dissolution/transformation of pre-existing phases and not associated with a poorly-

934 crystalline leached clay residue. The spectral characteristics of halloysite-jarosite

935 mixtures, as presented in Graff et al. [2012], are more consistent with the doublet

936 characteristics observed on Mars than are the spectral characteristics of a Si-rich

937 leached clay residue. Spectra of the latter tend to exhibit absorptions in the ~2.2-2.3

938 µm region that span a slightly broader wavelength range than the martian examples

939 and for which the individual absorptions are not as narrow or well-defined as the

940 martian examples. Halloysite and kaolinite are spectrally similar in the ~2.2-2.3 µm

941 region, thus halloysite-jarosite, kaolinite-jarosite, and montmorillonite/illite-

942 jarosite mixtures could all be consistent with the martian doublet materials.

943

944 Perhaps akin to the observations by Graff et al. [2012], laboratory

945 experiments by Tosca et al. [2008b] indicate that dissolution of basalt and addition

946 of sulfur could lead to fluids that form precipitates under oxidizing conditions

947 whose spectra also exhibit a doublet between 2.2 and 2.3 μm [see example spectrum

948 in Roach et al., 2010]. Though some of the precipitates generated during these

949 experiments were X-ray amorphous, others exhibited XRD patterns consistent with

950 neoformation of jarosite, Fe-oxides, and opaline silica [Tosca et al., 2008b]. VIS-NIR

42 951 reflectance spectra of these precipitates were also consistent with the presence of

952 opaline silica, minor jarosite, and an Al-OH bearing phase, though the mineralogical

953 host(s) of observed Al-OH bands at ~2.21 µm was not directly evident in the XRD

954 data and it/they may be below the XRD detection limit.

955

956

957 5.2 Formation Processes: Detrital and Authigenic

958

959 From a process point of view the distinction between the scenarios described

960 above may be somewhat semantic. The important point is simply the recognition

961 that dissolution and alteration of basaltic materials on Mars, including cases where

962 clay minerals may already be present, can lead to fluids capable of forming jarosite

963 and Al-rich phyllosilicates. However, co-occurrence in the martian deposits does not

964 necessarily imply that both jarosite and Al-phyllosilicates were formed at the same

965 time, under the same conditions, or even in the same region. Either mineral could be

966 detrital or authigenic in the doublet-bearing units, and detailed geologic context in

967 conjunction with an assessment of the full mineral assemblage is needed to assess

968 which scenario is most likely.

969

970 Given that halloysite, kaolinite, muscovite/illite, and montmorillonite all

971 exhibit an absorption feature at ~2.21 µm and that all have been identified on Mars

972 from CRISM data [Ehlmann et al., 2009], it may be difficult to unambiguously

973 determine which Al-bearing phyllosilicate is present in these purported mixtures

43 974 using VIS-NIR reflectance spectroscopy alone. There is evidence in some of the

975 martian spectra that the doublet feature may exhibit a shoulder on the short

976 wavelength side of the ~2.21 µm absorption, indicating a weak and narrow

977 absorption may be present in the ~2.15-2.21 µm region. If true, this could be more

978 consistent with the presence of kaolinite or halloysite as the Al-bearing

979 phyllosilicate as opposed to illite or muscovite or montmorillonite. The presence of

980 jarosite would indicate low-pH conditions (at least locally), and previous studies

981 have noted that kaolin minerals can be formed under acidic conditions [e.g., Fialips

982 et al., 2000]. On Mars, oxidation of fluids enriched in ferrous iron could lead to local

983 (and transient) acidification, which could promote jarosite formation if sulfides are

984 present or if sulfate is present in solution [e.g., Hurowitz et al., 2009, 2010].

985

986 The processes described above indicate the possibility that both the Al-

987 phyllosilicates and jarosite are authigenic and thus indicative of local aqueous

988 geochemical conditions. Though some Al-phyllosilicates can form under moderate

989 to low-pH conditions, particularly kaolinite, an alternative possibility is that the Al-

990 phyllosilicates in the martian doublet materials are detrital. In this scenario the

991 jarosite would be indicative of local aqueous conditions and the phyllosilicates

992 would reflect eolian or aqueous sediment transport processes. If true, the doublet-

993 bearing units in Valles Marineris may be similar to Rio Tinto in terms of mineralogy

994 and process, in which case the relative proportions of Al-phyllosilicate and jarosite

995 in a given deposit could reflect the relative contributions of sediment input and

996 mineral precipitation, respectively. Recent experimental work by Dixon et al. [2015]

44 997 demonstrated that jarosite dissolution is not strongly dependent on hydrologic

998 conditions and that 1 mm jarosite particles are unlikely to survive for more than 104

999 - 105 years even at low flow rates (<0.01 L hr-1). Therefore, while it is possible that

1000 the jarosite on Mars may be detrital, this is unlikely to be the case for long-lived

1001 fluvial transport systems, similar to the lack of jarosite observed in the older

1002 deposits at Rio Tinto. That stated, because we cannot constrain the particle size of

1003 the jarosite we cannot rule out eolian transport for this component.

1004

1005 A detailed ‘source to sink’ and stratigraphic study of the geologic setting of

1006 each of the doublet bearing units on Mars is needed to help determine which ones

1007 may represent diagenetic/authigenic phases (do observed mineral boundaries cut

1008 across the stratigraphy?) versus those that may be detrital (do variations in

1009 mineralogy conform to stratigraphic boundaries?), which is beyond the scope of the

1010 current study. However, given the diversity in morphologic properties of the

1011 doublet-bearing units within Valles Marineris and the results of the case study

1012 described above, it is likely that multiple formation processes are responsible even

1013 though the spectral signatures are similar.

1014

1015 The massive brecciated deposit in Ius Chasma that shows a relative

1016 transition from more jarosite to more Al-phyllosilicate within a single geologic unit

1017 (first example in our case study, Figure 11c) could be evidence of in situ alteration.

1018 Similarly, the thin draping units such as the one shown in Figure 10a are difficult to

1019 explain by processes other than airfall deposition [e.g., Weitz et al., 2015], which

45 1020 would suggest a relatively fine grain size and thus make the original sediment more

1021 susceptible to chemical alteration if it has not been welded. In this case, the draping

1022 unit may be an airfall dust or volcanic ash deposit, either of which could contain

1023 sulfur, that was later altered by water, possibly in the form of thins films from

1024 melting snow or ice. Atmospheric sources of sulfur (e.g., from volcanic emissions)

1025 could also contribute to authigenic sulfate formation. Regardless of the source, local

1026 differences in iron and/or sulfur content of the original sediment would lead to

1027 different authigenic mineral assemblages, such as different amounts of Al-

1028 phyllosilicate and jarosite [Michalski et al., 2013].

1029

1030 The second example in our case study of Ius Chasma , where Mg/Fe clay-

1031 bearing strata are unconformably overlain by layers have differing amounts of

1032 inferred jarosite and Al-phyllosilicate (Figures 11-12), may be more consistent with

1033 changes in sediment flux and/or source region. The stratigraphic order of these

1034 units indicates the large Fe/Mg clay unit pre-dates deposition of the units with the

1035 doublet feature. Indeed, there is no clear genetic relationship between the Mg/Fe

1036 clay unit and the doublet units, suggesting the latter are inconsistent with in situ

1037 alteration or leaching of the former. One possible interpretation is that the different

1038 proportions of the Al-phyllosilicate component, which correspond to distinct

1039 stratigraphic units, represent changes in sediment flux and or source region,

1040 perhaps where an increase in Al-phyllosilicate corresponds to a more highly

1041 weathered source region. Unlike the other example we presented for this region, the

1042 relative strength of the ~2.27 µm jarosite absorption also appears to conform to

46 1043 stratigraphic boundaries. This may indicate that the jarosite is also detrital, but an

1044 authigenic origin is possible as well since pore fluid flow or chemistry may be

1045 stratigraphically controlled.

1046

1047 If this sequence in Ius Chasma does in fact represent different relative

1048 amounts of authigenic jarosite and detrital Al-phyllosilicate, then we are potentially

1049 observing an upsection transition from precipitate (jarosite) dominated lithologies

1050 to ones dominated by clastic input. Such stratigraphic changes in mineralogy could

1051 be formed in a variety of environments, and without modal abundances for these

1052 minerals it is difficult to narrow the possible interpretations. However, possibilities

1053 include decreasing either the abundance of jarosite cement (which may or may not

1054 correspond to a change in absolute Al-phyllosilicate content), the availability of

1055 water, and/or Fe or S availability. Alternatively, a change in redox conditions, pH, or

1056 an increase in detrital input through time could also influence the relative

1057 proportions of jarosite and Al-phyllosilicate. Regardless, the observation that there

1058 are deposits suggestive of both in situ alteration and a combination of detrital influx

1059 of clay and precipitation of sulfate shows that multiple aqueous processes were

1060 likely responsible for formation of the doublet-bearing materials.

1061

1062 One of the most intriguing features of the doublet-bearing rocks at Rio Tinto

1063 is that the associated mineral assemblage is not preserved through geologic time in

1064 a terrestrial environment. Influx of fresh meteoritic water removes sulfates,

1065 including jarosite, and results in a net increase in ferric oxides, predominantly

47 1066 goethite, through time [Elwood Madden et al., 2004; Golden et al., 2005]. The

1067 transition from young sulfate-rich streambed deposits to older ferric oxide-rich

1068 terraces takes place in less than a million years, which is in agreement with jarosite

1069 dissolution rates reported by Dixon et al. [2015]. That study also noted that jarosite

1070 dissolution rate is more likely to be controlled by mineral-surface reaction rates

1071 rather than hydrologic conditions. Assuming that the Valles Marineris deposits have

1072 been exposed on the surface since at least the Early , the apparent

1073 persistence of jarosite indicates that interaction between the doublet materials and

1074 water (e.g., diagenetic overprinting) has been rather limited since the time of

1075 jarosite formation or deposition. Thus, while water is required for the formation of

1076 jarosite and phyllosilicates, and for some occurrences of the doublet materials water

1077 may have acted as a sediment transport agent, it is likely that post-deposition water-

1078 rock interaction has been significantly limited in these regions for hundreds of

1079 millions of years or longer.

1080

1081 In summary, it is important to keep in mind that the presence of a mineral or

1082 mineral assemblage in and of itself is not necessarily indicative of local aqueous

1083 geochemistry for a given outcrop or region. Mineral identifications, particularly

1084 those determined at spatial scales that are typical of orbital spectrometers (10s of

1085 meter per pixel or greater), must be placed in proper context to determine how they

1086 relate to geologic processes and environmental conditions. Specifically, it is critical

1087 to determine which minerals may be detrital versus authigenic (to the extent

1088 possible) and to assess the textural relationships between different mineral phases.

48 1089 Do sulfates and/or phyllosilicates occur as cementing agents, vein fill, or clasts? For

1090 minerals that are indicative of specific chemical conditions, such as jarosite, does the

1091 abundance and/or presence of the phases require those conditions to be persistent

1092 or might they represent short-lived exceptions to very different background

1093 conditions? As it stands, addressing such questions typically requires detailed small-

1094 scale observations that are only achievable by rovers or humans, thus caution

1095 should be exercised when interpreting orbital observations in terms of specific

1096 geologic processes, particularly at small spatial scales.

1097

1098 6. Summary and Conclusions

1099

1100 An enigmatic doublet feature in reflectance spectra has been observed

1101 alongside hydrated mineral deposits in a number of regions on Mars. Previous

1102 studies have proposed that this spectral signature could arise from mixtures of

1103 hydrated minerals with known absorptions between 2.2 and 2.3 µm, including

1104 jarosite, gypsum, silica, Fe/Mg smectite, and Al-phyllosilicates. Possible spectral and

1105 mineralogical analogs from Rio Tinto are analyzed with a combination of laboratory,

1106 field, and aerial reflectance spectra and XRD analysis of the mineralogy and reveal

1107 that the doublet arises in rocks containing both illite or muscovite and jarosite. The

1108 martian doublet materials are also consistent with a mixture of Al-phyllosilicate and

1109 jarosite because (1) the wavelength positions of the martian doublet absorptions

1110 are similar to those observed in Rio Tinto data collected at lab, field, and airborne

1111 spatial scales, (2) the wavelength positions and widths of individual absorptions are

49 1112 consistent with the presence of Al-OH and Fe-OH-bearing minerals, and (3) the

1113 strengths of the two absorptions in the doublet vary independently, implying the

1114 spectra represent a mixture of two or more components. Likely Al-phyllosilicates for

1115 the martian examples include montmorillonite, illite, and kaolinite/halloysite.

1116

1117 Two scenarios are proposed for the creation of these clay-sulfate

1118 assemblages on Mars in the context of a case study of deposits in Ius Chasma. In one

1119 example there are clear mineralogical changes (variations in relative proportions of

1120 jarosite and Al-phyllosilicate) within a single breccia unit. These deposits are

1121 consistent with in situ jarosite formation, possibly linked to diagenetic fluids. In a

1122 second example, observed changes in the relative proportions of the two minerals

1123 appear to follow stratigraphic boundaries, and in all cases the doublet-bearing units

1124 lie unconformably over Mg/Fe clay-bearing strata. A detrital origin for both phases

1125 cannot be ruled out in this case, but it is also plausible that the observed changes in

1126 mineralogy reflect changes in the relative contribution of precipitates (e.g., jarosite)

1127 versus detritus (e.g., Al-phyllosilicates) through time. Both detrital and authigenic

1128 origins are consistent for different doublet-bearing deposits in Ius Chasma, and the

1129 complexity of these deposits demonstrates the need for detailed mapping of each

1130 occurrence to decipher their likely geologic history. The results of such detailed

1131 mapping studies will ultimately provide the most insight into what these deposits

1132 record about aqueous activity, preservation potential, and water-rock-atmosphere

1133 interaction on Mars.

1134

50 1135 Acknowledgements

1136 The authors would like to acknowledge the support of the NASA Astrobiology

1137 Institute (MIT team, award NNA13AA90A) for funding this work. We also thank

1138 John Mustard for providing the HyMap data and the NASA Reflectance Experiment

1139 LABoratory (RELAB) facility for providing some of the spectral data used in this

1140 study. Comments from an anonymous reviewer and Liz Rampe helped to improve

1141 this manuscript and their efforts are graciously acknowledged. All data used in this

1142 paper are available upon request through the corresponding author

1143 ([email protected]).

1144 1145 References 1146 1147 Altheide, T.S., Chevrier, V.F., Noe Dobrea, E., 2010. Mineralogical characterization of acid weathered 1148 phyllosilicates with implications for secondary martian deposits. Geochimica et Cosmochimica 1149 Acta 74, 6232–6248. doi:10.1016/j.gca.2010.08.005 1150 Arvidson, R.E., Ruff, S.W., Morris, R.V., Ming, D.W., Crumpler, L.S., Yen, A.S., Squyres, S.W., Sullivan, 1151 R.J., Bell, J.F., Cabrol, N.A., Clark, B.C., Farrand, W.H., Gellert, R., Greenberger, R., Grant, J.A., 1152 Guinness, E.A., Herkenhoff, K.E., Hurowitz, J.A., Johnson, J.R., Klingelhöfer, G., Lewis, K.W., 1153 Li, R., McCoy, T.J., Moersch, J., McSween, H.Y., Murchie, S.L., , M., Schröder, C., 1154 Wang, A., Wiseman, S., Madsen, M.B., Goetz, W., McLennan, S.M., 2008. Mars Rover 1155 Mission to the , Gusev Crater: Mission overview and selected results from the 1156 Cumberland Ridge to . Journal of Geophysical Research 113. 1157 doi:10.1029/2008JE003183 1158 Bibring, J.-P., 2006. Global Mineralogical and Aqueous Mars History Derived from OMEGA/Mars 1159 Express Data. Science 312, 400–404. doi:10.1126/science.1122659 1160 Bibring, J.-P., Arvidson, R.E., Gendrin, A., Gondet, B., Langevin, Y., Le Mouelic, S., Mangold, N., Morris, 1161 R.V., Mustard, J.F., Poulet, F., Quantin, C., Sotin, C., 2007. Coupled Ferric Oxides and Sulfates 1162 on the . Science 317, 1206–1210. doi:10.1126/science.1144174 1163 Bish, D.L., Howard, S.A., 1988. Quantitative phase analysis using the Rietveld method. Journal of Applied 1164 Crystallography 21, 86–91. doi:10.1107/S0021889887009415 1165 Bishop, J.L., Loizeau, D., McKeown, N.K., Saper, L., Dyar, M.D., Des Marais, D.J., Parente, M., Murchie, 1166 S.L., 2013. What the ancient phyllosilicates at Mawrth Vallis can tell us about possible 1167 habitability on early Mars. Planetary and Space Science 86, 130–149. 1168 doi:10.1016/j.pss.2013.05.006 1169 Bristow, T.F., Bish, D.L., Vaniman, D.T., Morris, R.V., Blake, D.F., Grotzinger, J.P., Rampe, E.B., Crisp, 1170 J.A., Achilles, C.N., Ming, D.W., Ehlmann, B.L., King, P.L., Bridges, J.C., Eigenbrode, J.L., 1171 Sumner, D.Y., Chipera, S.J., Moorokian, J.M., Treiman, A.H., Morrison, S.M., Downs, R.T., 1172 Farmer, J.D., Marais, D.D., Sarrazin, P., Floyd, M.M., Mischna, M.A., McAdam, A.C., 2015. The 1173 origin and implications of clay minerals from Yellowknife Bay, Gale crater, Mars. American 1174 Mineralogist 100, 824–836. doi:10.2138/am-2015-5077CCBYNCND 1175 Burns, R.G., 1993. Rates and mechanisms of chemical weathering of ferromagnesian silicate minerals on 1176 Mars. Geochimica et Cosmochimica Acta 57, 4555–4574. doi:10.1016/0016-7037(93)90182-V

51 1177 Burns, R.G., 1987. Gossans on Mars. 18th Lunar and Planetary Science Conference 713–721. 1178 Carter, J., Poulet, F., Bibring, J.-P., Mangold, N., Murchie, S., 2013. Hydrous minerals on Mars as seen by 1179 the CRISM and OMEGA imaging spectrometers: Updated global view: HYDROUS MINERALS 1180 ON MARS: GLOBAL VIEW. Journal of Geophysical Research: Planets 118, 831–858. 1181 doi:10.1029/2012JE004145 1182 Cavanagh, P.D., Bish, D.L., Blake, D.F., Vaniman, D.T., Morris, R.V., Ming, D.W., Rampe, E.B., Achilles, 1183 C.N., Chipera, S.J., Treiman, A.H., Downs, R.T., Morrison, S.M., Fendrich, K.V., Yen, A.S., 1184 Grotzinger, J., Crisp, J.A., Bristow, T.F., Sarrazin, P., Farmer, J.D., Des Marais, Stolper, E., 1185 Morookian, J.M., Wilson, M.A., Spanovich, N., Anderson, R., 2015. Confidence Hills Mineralogy 1186 and CheMin Results from Base of Mt. Sharp, Pahrump Hills, Gale Crater, Mars. 46th Lunar and 1187 Planetary Science Conference Abstract #2735. 1188 Chemtob, S.M., Arvidson, R.E., FernáNdez-Remolar, D., Amils, R., 2006. Identification of Hydrated 1189 Sulfates Collected in the Northern Rio Tinto Valley by Reflectance and Raman Spectroscopy. 1190 Lunar and Planetary Science XXXVII Abstract 1941. 1191 Clark, R.N., Roush, T.L., 1984. Reflectance spectroscopy: Quantitative analysis techniques for remote 1192 sensing applications. Journal of Geophysical Research 89, 6329. doi:10.1029/JB089iB07p06329 1193 Dixon, E.M., Elwood Madden, A.S., Hausrath, E.M., Elwood Madden, M.E., 2015. Assessing 1194 hydrodynamic effects on jarosite dissolution rates, reaction products, and preservation on Mars: 1195 Hydrodynamics effects Mars Minerals. Journal of Geophysical Research: Planets 120, 625–642. 1196 doi:10.1002/2014JE004779 1197 Ehlmann, B.L., Mustard, J.F., Murchie, S.L., Bibring, J.-P., Meunier, A., Fraeman, A.A., Langevin, Y., 1198 2011. Subsurface water and clay mineral formation during the early history of Mars. Nature 479, 1199 53–60. doi:10.1038/nature10582 1200 Ehlmann, B.L., Mustard, J.F., Swayze, G.A., Clark, R.N., Bishop, J.L., Poulet, F., Des Marais, D.J., Roach, 1201 L.H., Milliken, R.E., Wray, J.J., Barnouin-Jha, O., Murchie, S.L., 2009. Identification of hydrated 1202 silicate minerals on Mars using MRO-CRISM: Geologic context near Nili Fossae and implications 1203 for aqueous alteration. Journal of Geophysical Research 114. doi:10.1029/2009JE003339 1204 Elwood Madden, M.E., Bodnar, R.J., Rimstidt, J.D., 2004. Jarosite as an indicator of water-limited 1205 chemical weathering on Mars. Nature 431, 821–823. doi:10.1038/nature02971 1206 Farrand, W.H., Glotch, T.D., Rice, J.W., Hurowitz, J.A., Swayze, G.A., 2009. Discovery of jarosite within 1207 the Mawrth Vallis region of Mars: Implications for the geologic history of the region. Icarus 204, 1208 478–488. doi:10.1016/j.icarus.2009.07.014 1209 Fernández-Remolar, D.C., 2003. Geological record of an acidic environment driven by iron 1210 hydrochemistry: The Tinto River system. Journal of Geophysical Research 108. 1211 doi:10.1029/2002JE001918 1212 Fernández-Remolar, D.C., Knoll, A.H., 2008. Fossilization potential of iron-bearing minerals in acidic 1213 environments of Rio Tinto, Spain: Implications for Mars exploration. Icarus 194, 72–85. 1214 doi:10.1016/j.icarus.2007.10.009 1215 Fernández-Remolar, D.C., Morris, R.V., Gruener, J.E., Amils, R., Knoll, A.H., 2005. The Río Tinto Basin, 1216 Spain: Mineralogy, sedimentary geobiology, and implications for interpretation of outcrop rocks at 1217 Meridiani Planum, Mars. Earth and Planetary Science Letters 240, 149–167. 1218 doi:10.1016/j.epsl.2005.09.043 1219 Fernández-Remolar, D.C., Prieto-Ballesteros, O., Gómez-Ortíz, D., Fernández-Sampedro, M., Sarrazin, P., 1220 Gailhanou, M., Amils, R., 2011. Río Tinto sedimentary mineral assemblages: A terrestrial 1221 perspective that suggests some formation pathways of phyllosilicates on Mars. Icarus 211, 114– 1222 138. doi:10.1016/j.icarus.2010.09.008 1223 Fialips, C.-I., Petit, S., Decarreau, A., Beaufort, D., 2000. Influence of synthesis pH on kaolinite 1224 “crystallinity” and surface properties. Clays and Clay Minerals 48, 173–184. 1225 Gendrin, A., 2005. Sulfates in Martian Layered Terrains: The OMEGA/Mars Express View. Science 307, 1226 1587–1591. doi:10.1126/science.1109087 1227 Golden, D.C., Ming, D.W., Morris, R.V., Mertzman, S.A., 2005. Laboratory-simulated acid-sulfate 1228 weathering of basaltic materials: Implications for formation of sulfates at Meridiani Planum and 1229 Gusev crater, Mars. Journal of Geophysical Research 110. doi:10.1029/2005JE002451 1230 Graff, T.G., Morris, R.V., Achilles, C.N., Agresti, A.G., Ming, D.W., Hamilton, J.C., Mertzman, S.A., 1231 , J., 2012. Chemical and Mineralogical Characterization of Acid-Sulfate Alteration of

52 1232 Basaltic Material on Mauna Kea Volcano, Hawaii: Jarosite and Hydrated Halloysite. Lunar and 1233 Planetary Science XLIII Abstract, 2639. 1234 Grotzinger, J.P., Arvidson, R.E., Bell, J.F., Calvin, W., Clark, B.C., Fike, D.A., Golombek, M., , 1235 R., Haldemann, A., Herkenhoff, K.E., Jolliff, B.L., Knoll, A.H., Malin, M., McLennan, S.M., 1236 Parker, T., Soderblom, L., Sohl-Dickstein, J.N., Squyres, S.W., Tosca, N.J., Watters, W.A., 2005. 1237 Stratigraphy and sedimentology of a dry to wet eolian depositional system, Burns formation, 1238 Meridiani Planum, Mars. Earth and Planetary Science Letters 240, 11–72. 1239 doi:10.1016/j.epsl.2005.09.039 1240 Grotzinger, J.P., Gupta, S., Malin, M.C., Rubin, D.M., Schieber, J., Siebach, K., Sumner, D.Y., Stack, 1241 K.M., Vasavada, A.R., Arvidson, R.E., Calef, F., Edgar, L., Fischer, W.F., Grant, J.A., Griffes, J., 1242 Kah, L.C., Lamb, M.P., Lewis, K.W., Mangold, N., Minitti, M.E., Palucis, M., Rice, M., 1243 Williams, R.M.E., Yingst, R.A., Blake, D., Blaney, D., Conrad, P., Crisp, J., Dietrich, W.E., 1244 Dromart, G., Edgett, K.S., Ewing, R.C., Gellert, R., Hurowitz, J.A., Kocurek, G., Mahaffy, P., 1245 McBride, M.J., McLennan, S.M., Mischna, M., Ming, D., Milliken, R., Newsom, H., Oehler, D., 1246 Parker, T.J., Vaniman, D., Wiens, R.C., Wilson, S.A., 2015. Deposition, exhumation, and 1247 paleoclimate of an ancient lake deposit, Gale crater, Mars. Science 350, aac7575–aac7575. 1248 doi:10.1126/science.aac7575 1249 Grotzinger, J.P. et al., 2014. A Habitable Fluvio-Lacustrine Environment at Yellowknife Bay, Gale Crater, 1250 Mars. Science 343, 1242777–1242777. doi:10.1126/science.1242777 1251 Hurowitz, J.A., Fischer, W.W., Tosca, N.J., Milliken, R.E., 2010. Origin of acidic surface waters and the 1252 evolution of atmospheric chemistry on early Mars. Nature Geoscience 3, 323–326. 1253 doi:10.1038/ngeo831 1254 Hurowitz, J.A., Tosca, N.J., Dyar, M.D., 2009. Acid production by FeSO4{middle dot}nH2O dissolution 1255 and implications for terrestrial and martian aquatic systems. American Mineralogist 94, 409–414. 1256 doi:10.2138/am.2009.3085 1257 Klingelhofer, G., 2004. Jarosite and Hematite at Meridiani Planum from Opportunity’s Mossbauer 1258 Spectrometer. Science 306, 1740–1745. doi:10.1126/science.1104653 1259 Langevin, Y., 2005. Sulfates in the North Polar Region of Mars Detected by OMEGA/Mars Express. 1260 Science 307, 1584–1586. doi:10.1126/science.1109091 1261 Lichtenberg, K.A., Arvidson, R.E., Morris, R.V., Murchie, S.L., Bishop, J.L., Fernandez Remolar, D., 1262 Glotch, T.D., Noe Dobrea, E., Mustard, J.F., Andrews-Hanna, J., Roach, L.H., 2010. Stratigraphy 1263 of hydrated sulfates in the sedimentary deposits of Aram Chaos, Mars. Journal of Geophysical 1264 Research 115. doi:10.1029/2009JE003353 1265 Madejová, J., Pentrák, M., Pálková, H., Komadel, P., 2009. Near-infrared spectroscopy: A powerful tool in 1266 studies of acid-treated clay minerals. Vibrational Spectroscopy 49, 211–218. 1267 doi:10.1016/j.vibspec.2008.08.001 1268 Metz, J., Grotzinger, J., Okubo, C., Milliken, R., 2010. Thin-skinned deformation of sedimentary rocks in 1269 Valles Marineris, Mars. Journal of Geophysical Research 115. doi:10.1029/2010JE003593 1270 Metz, J.M., Grotzinger, J.P., Mohrig, D., Milliken, R., Prather, B., Pirmez, C., McEwen, A.S., Weitz, C.M., 1271 2009. Sublacustrine depositional fans in southwest Melas Chasma. Journal of Geophysical 1272 Research 114. doi:10.1029/2009JE003365 1273 Michalski, J.R., Niles, P.B., Cuadros, J., Baldridge, A.M., 2013. Multiple working hypotheses for the 1274 formation of compositional stratigraphy on Mars: Insights from the Mawrth Vallis region. Icarus 226, 1275 816–840. doi:10.1016/j.icarus.2013.05.024 1276 Milliken, R.E., Swayze, G.A., Arvidson, R.E., Bishop, J.L., Clark, R.N., Ehlmann, B.L., Green, R.O., 1277 Grotzinger, J.P., Morris, R.V., Murchie, S.L., Mustard, J.F., Weitz, C., 2008. Opaline silica in 1278 young deposits on Mars. Geology 36, 847. doi:10.1130/G24967A.1 1279 Murchie, S., Arvidson, R., Bedini, P., Beisser, K., Bibring, J.-P., Bishop, J., Boldt, J., Cavender, P., Choo, 1280 T., Clancy, R.T., Darlington, E.H., Des Marais, D., Espiritu, R., Fort, D., Green, R., Guinness, E., 1281 Hayes, J., Hash, C., Heffernan, K., Hemmler, J., Heyler, G., Humm, D., Hutcheson, J., Izenberg, 1282 N., Lee, R., Lees, J., Lohr, D., Malaret, E., Martin, T., McGovern, J.A., McGuire, P., Morris, R., 1283 Mustard, J., Pelkey, S., Rhodes, E., Robinson, M., Roush, T., Schaefer, E., Seagrave, G., Seelos, 1284 F., Silverglate, P., Slavney, S., Smith, M., Shyong, W.-J., Strohbehn, K., Taylor, H., Thompson, 1285 P., Tossman, B., Wirzburger, M., Wolff, M., 2007. Compact Reconnaissance Imaging 1286 Spectrometer for Mars (CRISM) on Mars Reconnaissance Orbiter (MRO). Journal of Geophysical 1287 Research 112. doi:10.1029/2006JE002682

53 1288 Mustard, J.F., Murchie, S.L., Pelkey, S.M., Ehlmann, B.L., Milliken, R.E., Grant, J.A., Bibring, J.-P., 1289 Poulet, F., Bishop, J., Dobrea, E.N., Roach, L., Seelos, F., Arvidson, R.E., Wiseman, S., Green, 1290 R., Hash, C., Humm, D., Malaret, E., McGovern, J.A., Seelos, K., Clancy, T., Clark, R., Marais, 1291 D.D., Izenberg, N., Knudson, A., Langevin, Y., Martin, T., McGuire, P., Morris, R., Robinson, 1292 M., Roush, T., Smith, M., Swayze, G., Taylor, H., Titus, T., Wolff, M., 2008. Hydrated silicate 1293 minerals on Mars observed by the Mars Reconnaissance Orbiter CRISM instrument. Nature 454, 1294 305–309. doi:10.1038/nature07097 1295 Noe Dobrea, E., Michalski, J.R., Swayze, G.A., 2011. Aqueous Mineralogy and Stratigraphy at and Around 1296 the Proposed Mawrth Vallis MSL Landing Site: New Insights into the Aqueous History of the 1297 Region. The International Journal of Mars Science and Exploration 6, 32–46. 1298 Pelkey, S.M., Mustard, J.F., Murchie, S., Clancy, R.T., Wolff, M., Smith, M., Milliken, R., Bibring, J.-P., 1299 Gendrin, A., Poulet, F., Langevin, Y., Gondet, B., 2007. CRISM multispectral summary products: 1300 Parameterizing mineral diversity on Mars from reflectance. Journal of Geophysical Research 112. 1301 doi:10.1029/2006JE002831 1302 Poulet, F., Bibring, J.-P., Mustard, J.F., Gendrin, A., Mangold, N., Langevin, Y., Arvidson, R.E., Gondet, 1303 B., Gomez, C., Berthé, M., Bibring, J.-P., Langevin, Y., Erard, S., Forni, O., Gendrin, A., Gondet, 1304 B., Manaud, N., Poulet, F., Poulleau, G., Soufflot, A., Combes, M., Drossart, P., Encrenaz, T., 1305 Fouchet, T., Melchiorri, R., Bellucci, G., Altieri, F., Formisano, V., Fonti, S., Capaccioni, F., 1306 Cerroni, P., Coradini, A., Korablev, O., Kottsov, V., Ignatiev, N., Titov, D., Zasova, L., Mangold, 1307 N., Pinet, P., Schmitt, B., Sotin, C., Hauber, E., Hoffmann, H., Jaumann, R., Keller, U., Arvidson, 1308 R., Mustard, J., Forget, F., 2005. Phyllosilicates on Mars and implications for early martian 1309 climate. Nature 438, 623–627. doi:10.1038/nature04274 1310 Roach, L.H., Mustard, J., Gendrin, A., Fernández-Remolar, D., Amils, R., Amaral-Zettler, L., 2006. 1311 Finding mineralogically interesting targets for exploration from spatially coarse visible and near 1312 IR spectra. Earth and Planetary Science Letters 252, 201–214. doi:10.1016/j.epsl.2006.09.044 1313 Roach, L.H., Mustard, J.F., Swayze, G., Milliken, R.E., Bishop, J.L., Murchie, S.L., Lichtenberg, K., 2010. 1314 Hydrated mineral stratigraphy of Ius Chasma, Valles Marineris. Icarus 206, 253–268. 1315 doi:10.1016/j.icarus.2009.09.003 1316 Sobron, P., Bishop, J.L., Blake, D.F., Chen, B., Rull, F., 2014. Natural Fe-bearing oxides and sulfates from 1317 the Rio Tinto Mars analog site: Critical assessment of VNIR reflectance spectroscopy, laser 1318 Raman spectroscopy, and XRD as mineral identification tools. American Mineralogist 99, 1199– 1319 1205. doi:10.2138/am.2014.4595 1320 Sobron, P., Sanz, A., Acosta, T., Rull, F., 2009. A Raman spectral study of stream waters and efflorescent 1321 salts in Rio Tinto, Spain. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 1322 71, 1678–1682. doi:10.1016/j.saa.2008.06.035 1323 Squyres, S.W., 2004. In Situ Evidence for an Ancient Aqueous Environment at Meridiani Planum, Mars. 1324 Science 306, 1709–1714. doi:10.1126/science.1104559 1325 Sunshine, J.M., Pieters, C.M., 1993. Estimating Modal Abundances From the Spectra of Natural and 1326 Laboratory Pyroxene Mixtures Using the Modified Gaussian Model. Journal of Geophysical 1327 Research 98, 9075–9087. 1328 Thollot, P., Mangold, N., Ansan, V., Le Mouélic, S., Milliken, R.E., Bishop, J.L., Weitz, C.M., Roach, 1329 L.H., Mustard, J.F., Murchie, S.L., 2012. Most Mars minerals in a nutshell: Various alteration 1330 phases formed in a single environment in Noctis Labyrinthus: ALTERATION PHASES IN 1331 NOCTIS LABYRINTHUS. Journal of Geophysical Research: Planets 117, n/a–n/a. 1332 doi:10.1029/2011JE004028 1333 Tosca, N.J., 2004. Acid-sulfate weathering of synthetic Martian basalt: The acid fog model revisited. 1334 Journal of Geophysical Research 109. doi:10.1029/2003JE002218 1335 Tosca, N.J., Knoll, A.H., McLennan, S.M., 2008a. Water Activity and the Challenge for Life on Early 1336 Mars. Science 320, 1204–1207. doi:10.1126/science.1155432 1337 Tosca, N.J., McLennan, S.M., 2006. Chemical divides and evaporite assemblages on Mars. Earth and 1338 Planetary Science Letters 241, 21–31. doi:10.1016/j.epsl.2005.10.021 1339 Tosca, N.J., Milliken, R.E., Michel, F.M., 2008b. Smectite Formation on Early Mars: Experimental 1340 Constraints. Martian Phyllosilicates: Recorders of Aqueous Processes Abstract, 7030. 1341 Vaniman, D.T. et al., 2014. Mineralogy of a Mudstone at Yellowknife Bay, Gale Crater, Mars. Science 1342 343, 1243480–1243480. doi:10.1126/science.1243480

54 1343 Wang, A., Korotev, R.L., Jolliff, B.L., Haskin, L.A., Crumpler, L., Farrand, W.H., Herkenhoff, K.E., de 1344 Souza, P., Kusack, A.G., Hurowitz, J.A., Tosca, N.J., 2006. Evidence of phyllosilicates in Wooly Figure1345 1 Patch, an altered rock encountered at West Spur, Columbia Hills, by the Spirit rover in Gusev 1346 crater, Mars. Journal of Geophysical Research 111. doi:10.1029/2005JE002516 1347 Weitz, C.M., Bishop, J.L., Thollot, P., Mangold, N., Roach, L.H., 2011. Diverse mineralogies in two 1348 troughs of Noctis Labyrinthus, Mars. Geology 39, 899–902. doi:10.1130/G32045.1 1349 Weitz, C.M., Noe Dobrea, E., Wray, J.J., 2014. Mixtures of clays and sulfates within deposits in western 1350 Melas Chasma, Mars. Icarus. doi:10.1016/j.icarus.2014.04.009 1351 1352 1353 Figures

RELAB spectrum: Muscovite (C1XT09), Montmorillonite (CJS125), Gypsum (C1CC35), Nontronite (C1CY20), Opal (C1RM34). USGS spectrum: Jarosite (GDS98K) 1354 1355 Figure 1: Reflectance spectra of minerals with absorptions between 2.2 and 2.3 μm 1356 (source: RELAB database) are compared to CRISM spectra showing the ‘doublet’ 1357 absorption. Continuum-removed spectra from 2.1 to 2.35 μm (right) are shown to 1358 highlight absorption features in the wavelength region where the doublet occurs. 1359 None of the pure mineral spectra have absorption positions and depths that are a 1360 direct match to the doublet absorptions, thus a mineral mixture is implied. 1361

55 Spectral' Reitveld' Sample'ID Location Terrace Doublet Refinement RT1 Berrocal Modern RT2 Berrocal Modern RT3 Berrocal Modern RT41(mud,1loose) Berrocal Modern RT5 Anabel's1Garden,1Upper1Road N/A RT6 Anabel's1Garden,1Upper1Road N/A RT7 Anabel's1Garden,1Upper1Road N/A Yes RT8 Anabel's1Garden,1Upper1Road N/A RT9 Anabel's1Garden,1Upper1Road N/A Yes Yes RT10 Anabel's1Garden,1Upper1Road N/A Yes Yes RT11 Anabel's1Garden,1Upper1Road N/A Yes Yes RT12 Barranco1de1los1Locos Lower1Terrace Yes Yes RT13 Barranco1de1los1Locos Lower1Terrace RT14 Barranco1de1los1Locos Lower1Terrace RT15 Barranco1de1los1Locos Lower1Terrace Yes Yes RT16 Anabel's1Garden,1Stream11 Modern Yes Yes RT171(precipitate) Anabel's1Garden,1Stream12 Modern RT181(precipitate) Anabel's1Garden,1Stream12 Modern RT19 Anabel's1Garden,1Stream12 Modern Yes Yes RT201(precipitate) Anabel's1Garden,1Stream12 Modern RT21 Anabel's1Garden,1Stream12 Modern RT22 Anabel's1Garden,1Stream12 Modern RT23 Anabel's1Garden,1Stream12 Modern Yes RT24 Source1Spring Modern RT25 Source1Spring Modern RT26 Source1Spring Modern Yes Yes RT271(precipitate) Source1Spring Modern RT28 Alta1de1la1Mesa Upper1Terrace RT29 Alta1de1la1Mesa Upper1Terrace RT30 Alta1de1la1Mesa Upper1Terrace Yes RT31 Alta1de1la1Mesa Upper1Terrace RT32 Gossan1outcrop N/A RT33 Gossan1outcrop N/A 1362 RT34 Gossan1outcrop N/A 1363 1364 Table 1: A full list of samples collected at the modern streambed, lower and upper 1365 terraces at Rio Tinto. All samples are rocks unless otherwise noted. Each sample is 1366 measured with the ASD, FTIR, and XRD in the laboratory. Samples with a spectral 1367 doublet and those that have been analyzed with a Rietveld refinement are noted. 1368

56 1369 1370 Figure 2: Terraces of the Rio Tinto in the schematic diagram from Fernandez- 1371 Remolar et al. (2005) are matched with locations in the HyMap aerial hyperspectral 1372 image of the region that were sampled in this study. 1373

57 1374

1375 1376 Figure 3: A principal components analysis has been carried out on the same HyMap 1377 image as seen in Figure 2, and like colors in the image correspond to regions of 1378 spectral similarity (R = PCA band 7, G = band 3, B = band 5). The spectral doublet is 1379 associated with units that appear in a dark red color in this false-color image, which 1380 is particularly evident at Anabel’s Garden (inset) where a number of the doublet- 1381 bearing samples were collected. Variation in the spectral doublet from 2.2 – 2.3 μm 1382 is highlighted with spectra collected from a 500 m HyMap transect (black arrow). 1383 1384 Sample Rietveld4Results XRD4Peak4Areas Jarosite4 Muscovite4 Illite4 Peak4area:4 Peak4area:4 Sample Location Doublet? R (wt%) (wt%) (wt%) wp Musc/Ill Jarosite RT9$ AG$Mining$Terrace Yes 3 16 14 5.61 1717.8 639 RT10 AG$Mining$Terrace Yes 7 46 0 5.01 2438.3 673.4 RT11 AG$Mining$Terrace Yes 4 24 0 5.52 1239.4 573.1 RT12$ Barranco$de$los$Locos Yes 11 13 0 1.12 2115.3 1956.1 RT15$ Barranco$de$los$Locos Yes 37 12 0 1.21 2089.9 2050.9 RT16 Anabel's$Garden$(Stream) Yes 3 5 0 10.08 555.4 353.4 RT19 Anabel's$Garden$(stream) Yes 3 0 19 3.16 536.3 464.9 RT26 Source$Spring Yes 8 50 0 2.52 4655.9 2959.1 RT30 Alta$de$la$Mesa No 0 0 0 3.67 0 0 RT23 Anabel's$Garden$(Stream) No 32 0 0 1.66 0 1003.8 1385 RT7 Anabel's$Garden No 0 0 4 1.03 432.1 0

58 1386 Table 2: Sample IDs and origin of samples that were analyzed using XRD and 1387 reflectance spectroscopy. The jarosite and Al-clay (muscovite/illite) content of each 1388 sample was estimated using Rietveld refinemenFigure 4 t on the XRD patterns. 1389 1390

Band Centers 2000

1500

1000

Noise Counts 0.015 500 0 2.1 2.2 2.3 0.01 Wavelength of Band Center

Noise 0.005 2500 2000 0 1500 Counts 1000

500

1 km 0 0 0.005 0.01 0.015 Standard Deviation (from 1.7 - 1.85 μm)

1391 1392 Figure 4: Two Gaussian fit examples are shown for FRT00009B27 from Ius Chasma, 1393 Mars. Due to noise in the spectrum, the Gaussian mean may be a better 1394 representative of band depth minimum than the true reflectance value. Noise and 1395 spectral smile are further analyzed for a central subset of the image. Noise is 1396 calculated as the standard deviation of a flat, absorption free portion of the 1397 spectrum from 1.7 – 1.85 µm. An image and histogram of this standard deviation 1398 show that noise is typically < 0.005 in reflectance. Band center histograms are also 1399 shown for the central image subset, and the reduced spread in band centers 1400 compared to the full image suggests spectral smile plays a role in the band center 1401 position. 1402

59 Figure 5 1403

A B

C

1404 1405 Figure 5: Example spectra from the field and laboratory for the A) modernly 1406 precipitating streambed at Rio Tinto (source region, which includes Source Spring, 1407 Richi Spring, and Anabel’s Garden) and B) the ~2myr old terrace at Alto de la Mesa. 1408 These spectra show the transition from very hydrated, sulfate minerals to ferric 1409 oxides through time (with some clays possibly associated with mining activity). All 1410 of the mineralogies identified in hand samples are also found in the field spectra, but 1411 only a limited number of each is shown for the sake of clarity. In C), absorptions in 1412 powdered samples that show a doublet are specifically compared to jarosite 1413 spectral endmembers. Again, a very similar doublet signature is seen in both field 1414 and hand samples, but is not shown for clarity. Endmember mineral spectra from 1415 RELAB and USGS are shown for reference (RELAB: RM-JFM-036, RM-JFM-035, CC- 1416 JFM-013A, JB-JLB-752, JB-JLB-753, JB-JLB-757, JB-CMP-047, JB-CMP-054, CC-JFM-

60 1417 017-B, JB-JLB_256, CC-JFM-035, CL-TXH-014, JB-CMP-25-B; USGS: BZ93-1, GDS2 Figure 5 1418 Zeolite W1R1Bb) 1419 Jarosite (%)/Al-clay (%) = 0.17 Sample Fit Error

Jarosite (%)/Al-clay (%) = 0.74

Jarosite (%)/Al-clay (%) = 3.15

Goethite Muscovite/ Jarosite Quartz 1420 Illite 1421 Figure 6: XRD patterns with corresponding reflectance spectra from three 1422 powdered samples with different proportions of jarosite and Al-clay (RT 11, RT15, 1423 RT16). Major phases are labeled for the XRD patterns and absorptions 1424 corresponding to the doublet feature are marked by red arrows in the reflectance 1425 spectra plots. 1426

61 RT19 RT10

Figure 8 1427 1428 Figure 7: Example reflectance spectra acquired in the field and for hand sample, 1429 chip, and powder versions of two samples from different locations within the study 1430 region. Chips are <9 mm in diameter and powders were sieved to <45 µm. Spectra of 1431 these samples highlight the variation in absolute reflectance and band depth of the 1432 two absorptions in the doublet feature when measured at different spatial scales. 1433 Hand sample RT10 has a varnished surface for which the reflectance spectra exhibit 1434 a clear doublet feature, whereas spectra from the interior portions exhibit a much 1435 weaker long-wavelength (jarosite) absorption. 1436

Relative Band Depth vs. Al-Clay vs. Band Depth 1 1 XRD Peak Area 100

RT10 2 2 0.8 R = 0.75236 80 R = 0.48937 RT11 0.6 60

RT9 0.4 40 Al-clay (wt%) 0.2 20

0 0 0 0.2 0.4 0.6 0.8 1 0 0.02 0.04 0.06 0.08 RT26 Al−clay / (Al−clay+Jarosite) Band Depth at 2.205 microns Relative Band Depth vs. Jarosite vs. Band Depth 2 1 XRD Reitveld Refinement RT16 100 R2 = 0.75446 0.8 2 RT19 80 R = 0.85938

RT12 0.6 60 RT15 0.4 40

0.2 Jarosite (wt%) 20

0 0 0 0.2 0.4 0.6 0.8 1 0 0.02 0.04 0.06 0.08 Al−clay / (Al−clay+Jarosite) Band Depth at 2.269 microns Band Depth 1/ (Band Depth 1 + 2) 1437

62 1438 Figure 8: Comparison of XRD results with reflectance spectra (band depths) to 1439 show relationship between mineral abundances and the absorption feature 1440 strength. A direct relationship between relative amounts of jarosite and Al-clay and 1441 the strengths of the absorption features indicates the shorter wavelength 1442 absorption (band 1; ~2.21 µm) arises from Al-clay and the longer wavelength 1443 absorption (band 2; ~2.265) arises from the presence of jarosite. This suggests that 1444 relative band strengths are directly related to the relative proportions of the 1445 phyllosilicate and sulfate components. 1446 Image Band)1)(µm) 2σ)(µm) Band)2)(µm) 2σ)(µm) Mars FRT00009B27)full)image 2.2109 0.0264 2.2709 0.019 FRT00009B27)center 2.2006 0.0134 2.265 0.0112 FRT0000A396)full)image 2.2048 0.0154 2.2696 0.0132 FRT0000A396)center 2.2055 0.0134 2.2705 0.0136 FRT0000AD3D)full)image 2.2055 0.0188 2.2662 0.0156 FRT0000AD3D)center 2.2085 0.0142 2.2671 0.0108 Rio)Tinto 1447 HyMap 2.2025 0.013 2.2602 0.011 1448 Table 3: The results of Gaussian fits for every pixel in three CRISM images (Mars) 1449 and the HyMap image (Rio Tinto) that includes the mean band centers and the 2σ 1450 deviations on these band centers. A central subset of the CRISM images is also 1451 analyzed and reported results show a shift in band centers and a decrease in 1452 standard deviation.

63 1453 Figure 9

A 50 km FRT00009B27 FRT0000A396 FRT0000AD3D

5 km 5 km 5 km

B

4 4 x10 Band Centers x10 Band Centers 4 Band Centers C 6 8 9x10 5 7 8 6 7 4 5 6 3 4 5

Counts 4 Counts

Counts 3 2 3 2 1 2 1 1 0 0 0 2.15 2.2 2.25 2. 3 2.15 2.2 2.25 2. 3 2.15 2.2 2.25 2.3 Wavelength of Band Center Wavelength of Band Center Wavelength of Band Center

Relative Amplitude Relative Amplitude Relative Amplitude 4 4 x104 of Absorptions x10 of Absorptions x10 of Absorptions D 2 2.5 2.5 Jarosite Al-Clay 2 Rich Rich 1 Counts

1 Counts 1 Counts 0 0 0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5 Band 1 Amp./ Band 2 Amp. Band 1 Amp./ Band 2 Amp. Band 1 Amp./ Band 2 Amp. 1454 1455 Figure 9: Examples of doublet-bearing materials on Mars as observed in MRO 1456 CRISM images. (A) MOLA hillshade image showing the location of three study 1457 regions in Ius and Melas Chasma within the Valles Marineris. Zoom-in regions below 1458 show locations of CRISM images FRT00009B27, FRT0000A396, and FRT0000AD3D 1459 in context with nearby CRISM images (overlain on HiRISE and CTX images). 1460 Doublet-bearing units in CRISM images 9B27 and AD3D are light toned and tend to 1461 drape pre-existing topography; CRISM image A396 shows that the doublet-bearing

64 1462 material corresponds to a blocky, brecciated unit. (B) Examples of spectral variation 1463 within the doublet-bearing units, ranging from a stronger 2.205 μm absorption to a 1464 stronger 2.269 μm absorption. (C) Histograms showing positions of the 2.205 μm 1465 (blue) and 2.269 μm (red) band positions for every doublet-bearing pixel in each of 1466 the three images and (D) histograms showing the relative strength of the two 1467 absorption features (Band 1 = ~2.205 µm; Band 2 = ~2.269 µm). Values in (C) and 1468 (D) are based on Gaussian fits as described in the text. Data for each column in (B), 1469 (C), and (D) correspond to the CRISM FRT images labeled and highlighted in red in 1470 (A). 1471

1472 1473 Figure 10: Examples of doublet-bearing units as observed in MRO HiRISE images. 1474 These units are typically light-toned, contorted and/or blocky in nature. (A) 1475 Example of a doublet-bearing unit in Melas Chasma that drapes pre-existing 1476 topography and spans a wide range in elevation. (B) An example of doublet-bearing, 1477 light-toned units that may have slumped off the canyon walls and are now part of a 1478 brecciated unit on the floor. The morphology and spectral characteristics of these 1479 and similar deposits are described in detail in Roach et al. (2010), Metz et al. (2010), 1480 and Weitz et al. (2014). (HiRISE images: (A) PSP_007074_1725, (B) 1481 PSP_007509_1720).

65 1482

1483 1484 Figure 11: Relative amplitude of absorptions (band depth at 2.207/ band depth at 1485 2.269) in Ius Chasma and associated morphology. (A) False-color CRISM images (R – 1486 band 253, B – band 53, G = band 17, top) overlain on a CTX mosaic. Fe/Mg-clay 1487 bearing regions appear greenish in color and the doublet-bearing regions appear 1488 light-toned or white. (B) CRISM pixels with the doublet feature are highlighted by 1489 different colors that map the relative band strength of the two absorptions, with a 1490 gradient from more jarosite rich (red), to roughly equal jarosite/Al-phyllosilicate 1491 strength, to slightly more Al-phyllosilicate rich (blue), to very Al-phyllosilicate rich 1492 (yellow). (C) Close-up view of HiRISE image ESP_041121_1725 showing that there is 1493 no clear stratigraphic boundary associated with the transition from a more jarosite-

66 1494 rich “red” region to a more Al-phyllosilicate rich “blue” area in this location. (D) 1495 Close-up of HiRISE image PSP_007074_1725 showing an example where 1496 morphologically and stratigraphically distinct units correspond to the different 1497 relative band strengths. Here, white arrows indicate a clear stratigraphic boundary 1498 that also corresponds to a clear boundary between green and blue pixels seen in (B) 1499 (HiRISE images: ESP_041121_1725, PSP_007074_1725, CRISM images listed above). 1500

1501 1502 Figure 12: Evidence of the stratigraphic contacts between different doublet-bearing 1503 units that are identified as more jarosite- or Al-phyllosilicate-rich in the color map in 1504 Figure 11. (A) Context map from HiRISE image ESP_041121_1725 showing locations 1505 where three contacts are visible. (B) Shows the clay bearing unit (dark-toned, 1506 stratified with tilted bedding) unconformably overlain by the light-toned doublet 1507 unit, which has relative band strengths between 0.9 and 1.2 (i.e. the strength of the 1508 two absorptions in the doublet are similar). Next, in (C) there is a contact between 1509 this same light-toned unit and a slightly darker, more Al-phyllosilicate-rich unit, 1510 which appears to lie stratigraphically above it. Finally, (D) shows the contact 1511 between the brightest unit, which outcrops in smaller areas throughout Ius, and the

67 1512 light-toned doublet-bearing unit shown in (B) and (C). This bright unit is mapped as 1513 being more jarosite rich in Figure 8 and appears to be stratigraphically lower than 1514 the light-toned unit. It is not possible to determine the exact relationship between 1515 this bright material and the dark, stratified clay deposit in this location, but in other 1516 regions it appears as if the brightest (jarosite-rich) material is stratigraphically 1517 above the clay-bearing strata. Therefore, the increasing stratigraphic order appears 1518 to be: dark (clay-bearing) strata > brightest unit > light-toned unit > moderate- 1519 toned unit. The corresponding mineralogy is consistent with: clay-rich > more 1520 jarosite-rich > equal Al-phyllosilicate and jarosite > more Al-phyllosilicate rich (that 1521 is, within the doublet-bearing units the proportion of jarosite decreases upsection). 1522 1523 1524 1525 1526 1527 1528 1529 1530 1531 1532 1533 1534 Supplementary Material 1535 1536 1537 1538 1. Methods 1539 1540 Additional information from the methods section can be found here. 1541 1542 1543 1.1. CRISM image processing and Gaussian fitting 1544 1545 CRISM image cubes were corrected for atmospheric gases using the ‘volcano

1546 scan’ method described by Langevin et al. [2005] and Mustard et al. [2008]. All

1547 CRISM processing steps, including map-projection, were carried out with the freely

1548 available IDL CRISM Analysis Tools (CAT) [Pelkey et al., 2007] and the commercially

1549 available ENVI software package.

1550 CRISM spectra in regions of interest (ROIs) spanning ~100 to 1000 pixels

1551 were averaged to understand large-scale regional mineral assemblages. A spectrally

68 1552 bland or neutral ROI was chosen to represent background martian surface

1553 reflectance properties (e.g. martian dust), and every pixel within an image was

1554 divided by the spectrum corresponding to the neutral ROI within that same image.

1555 This method produces spectral ratios that highlight and accentuate reflectance

1556 properties of ROIs relative to those associated with typical martian dust and/or soil.

1557 Starting with the continuum-removed CRISM images, pixels whose spectra

1558 exceeded a user-defined (and image-dependent) band depth threshold at ~2.205

1559 μm were selected for Gaussian fitting. The threshold used here is band depth

1560 ≥0.005. Though it is not suggested that the same value can be used for every CRISM

1561 image due to differences in noise, for the images used in this study this value is

1562 larger than the background ‘noise’ in the ratio spectra. This was determined by

1563 calculating the mean and standard deviation over the 1.70-1.85 µm wavelength

1564 range (24 channels) for each ratio spectrum (pixel) in each CRISM image. This

1565 wavelength range was chosen because it corresponds to a ‘flat’ portion of the

1566 spectrum for materials in these CRISM images (i.e., there are no strong absorption

1567 features within this range).

1568 Gaussian mean (band center), standard deviation (band width), and

1569 amplitude (band strength) are allowed to vary within user-defined ranges. The

1570 Gaussian fits for at least 10 individual pixels are inspected for each image to ensure

1571 that the model fits are appropriate and residuals values for the spectral fits are

1572 reported for every pixel.

1573

1574 1.2 Field Observations

69 1575 The ASD FieldSpec 3 collects reflected radiation through a fiber optic cable

1576 that is connected to three detectors within the spectrometer housing. The bare fiber

1577 optic cable has a field of view (FOV) of ~25° but can be equipped with different

1578 foreoptics to reduce the FOV to as small as 1°. Field spectra measured at Rio Tinto

1579 using solar radiation were acquired using a 5° FOV foreoptic with a typical standoff

1580 distance of ~1 m from the target of interest. The contact probe uses an internal

1581 quartz-tungsten-halogen (QTH) bulb for illumination and measures a spot size of

1582 ~10 mm. The contact probe has the advantage that viewing/illumination geometry

1583 is fixed and thus identical for all measurements.

1584

1585 1.3 Laboratory Reflectance Measurements

1586 The light source for the laboratory reflectance measurements with the ASD

1587 FieldSpec consisted of an enclosed QTH bulb in combination with a fiber optic cable

1588 that has a 25° FOV. The source fiber was positioned to have a central incidence angle

1589 of 30° (from normal) and the receiving fiber (attached to the spectrometer) was

1590 positioned to have a central emergence angle of 0°. The heights of the fibers were

1591 adjusted as needed in order to fully illuminate and capture the entire sample in the

1592 field of view.

1593

1594 1.4 XRD Measurements and Rietveld Modeling

1595 The D2 Phaser is equipped with a low voltage Cu source and a Lynxeye 1-D

1596 detector. Measurements were made using a 0.6mm divergence slit, 3 mm

1597 antiscatter shield, 2.5˚soller slit and Ni filter and rotating stage (15 rpm).

70 1598 The Rietveld modeling incorporated instrumental parameters included the

1599 Cukα5_Berger emission profile, a 3rd order Chebyshev polynomial (background

1600 modeling), and specimen displacement correction. Strain and crystallite size

1601 broadening were varied to help fit the peak breadths while unit cell parameters and

1602 cation occupancies within each crystal structure were refined to model the

1603 appropriate peak position. A spherical harmonics correction was also applied to

1604 correct for any intensity variations due to preferred orientation during sample

1605 mounting.

Mineral Source Chlorite Zanazzi-et-al.-(2007) Goethite Yang-et-al.-(2006) Gypsum Boeyens-and-Ichharam-(2002) Illite Gaultieri-(2000) Jarosite Basciano-and-Peterson-(2007,-2008);-Scarlett-et-al.-(2010) Muscovite Birgatti-et-al.-(1998);-Birle-and-Tettenhorst-(1968) Pyrite Bayliss-(1977) 1606 Quartz Levien-et-al.-(1980) 1607 S Table 1: Structure file references for the Rietveld refinement. 1608 References: 1609 Basciano, L.C., Peterson, R.C., 2008. Crystal chemistry of the natrojarosite-jarosite and natrojarosite- 1610 hydronium jarosite solid-solution series: A synthetic study with full Fe site occupancy. American 1611 Mineralogist 93, 853–862. doi:10.2138/am.2008.2731 1612 Basciano, L.C., Peterson, R.C., 2007. Jarosite hydronium jarosite solid-solution series with full iron site 1613 occupancy: Mineralogy and crystal chemistry. American Mineralogist 92, 1464–1473. 1614 doi:10.2138/am.2007.2432 1615 Bayliss, P., 1977. Crystal structure refinement of a weakly anisotropic pyrite. American Mineralogist 62, 1616 1168–1172. 1617 Birle, J.D., 1968. Refined Muscovite Structure. Mineralogical Magazine 36, 883–886. 1618 doi:10.1180/minmag.1968.036.282.24 1619 Boeyens, J.C.A., Ichharam, V.V.H., 2002. Redetermination of the crystal structure of calcium sulphate 1620 dihydrate, CaSO4 · 2H2O. Zeitschrift für Kristallographie - New Crystal Structures 217. 1621 doi:10.1524/ncrs.2002.217.jg.9 1622 Brigatti, M.F., Frigieri, P., Luciano, P., 1998. Crystal chemistry of Mg-, Fe-bearing muscovites 2M1. 1623 American Mineralogist 83, 775–785. 1624 Gualtieri, A.F., 2000. Accuracy of XRPD QPA using the combined Rietveld–RIR method. Journal of 1625 Applied Crystallography 33, 267–278. doi:10.1107/S002188989901643X 1626 Levien, L., Prewitt, C.T., Weidner, D.J., 1980. Structure and elastic properties of quartz at pressure. 1627 American Mineralogist 65, 920–930.

71 1628 Ransom, B., Bennett, R.H., Baerwald, R., Shea, K., 1997. TEM study of in situ organic matter on 1629 continental margins: occurrence and the “monolayer” hypothesis. Marine Geology 138, 1–9. 1630 doi:10.1016/S0025-3227(97)00012-1 1631 Scarlett, N.V.Y., Grey, I.E., Brand, H.E.A., 2010. Ordering of iron vacancies in monoclinic jarosites. 1632 American Mineralogist 95, 1590–1593. doi:10.2138/am.2010.3591 1633 Yang, H., Lu, R., Downs, R.T., Costin, G., 2006. Goethite, α-FeO(OH), from single-crystal data. Acta 1634 Crystallographica Section E Structure Reports Online 62, i250–i252. 1635 doi:10.1107/S1600536806047258 1636 Zanazzi, P.F., Montagnoli, M., Nazzareni, S., Comodi, P., 2007. Structural effects of pressure on 1637 monoclinic chlorite: A single-crystal study. American Mineralogist 92, 655–661. 1638 doi:10.2138/am.2007.2341 1639

72